Academia.eduAcademia.edu

Kuby Immunology 7th Edition 2013CSIR NET LIFE SCIENCE

KUBY Immunology Icons Used in This Book Antigenic peptide T cell receptor Antibody CD3 Immature thymocyte TH cell CD8 TC cell Plasma cell B cell CD4 Class I MHC Cytokine Class II MHC Cytokine receptor Cytotoxic T cell Bone marrow stromal cell Neutrophil Basophil Eosinophil Dendritic cell Monocyte Macrophage Class I MHC Erythrocyte Antigen-presenting cell CD4 Altered self cell B cell Platelets Mast cell Class II MHC CD8 TC cell Natural killer cell TH cell KUBY Immunology Judith A. Owen Haverford College Jenni Punt Haverford College Sharon A. Stranford Mount Holyoke College with contributions by Patricia P. Jones Stanford University Seventh Edition W. H. Freeman and Company • New York Publisher: Susan Winslow Senior Acquisitions Editor: Lauren Schultz Associate Director of Marketing: Debbie Clare Marketing Assistant: Lindsay Neff Developmental Editor: Erica Champion Developmental Editor: Irene Pech Developmental Coordinator: Sara Ruth Blake Associate Media Editor: Allison Michael Library of Congress Control Number: 2012950797 North American Edition Cover image: ©2009 Pflicke and Sixt. Originally published in The Journal of Experimental Medicine. 206:2925-2935. doi:10.1084/jem.20091739. Image provided by Holger Pflicke and Michael Sixt. International Edition Cover design: Dirk Kaufman Cover image: Nastco/iStockphoto.com Supplements Editor: Yassamine Ebadat Senior Project Manager at Aptara: Sherrill Redd Photo Editor: Christine Buese Photo Researcher: Elyse Reider Art Director: Diana Blume Text Designer: Marsha Cohen Illustrations: Imagineering Illustration Coordinator: Janice Donnola North American Edition ISBN-13: 978-14292-1919-8 ISBN-10: 1-4292-1919-X International Edition ISBN-13: 978-14641-3784-6 ISBN-10: 1-4641-3784-6 © 1992, 1994, 1997, 2000, 2003, 2007, 2013 by W. H. Freeman and Company All rights reserved Production Coordinator: Lawrence Guerra Composition: Aptara®, Inc. Printing and Binding: RR Donnelley Printed in the United States of America First printing North American Edition W. H. Freeman and Company 41 Madison Avenue New York, NY 10010 www.whfreeman.com International Edition Macmillan Higher Education Houndmills, Basingstoke RG21 6XS, England www.macmillanhighered.com/international To all our students, fellows, and colleagues who have made our careers in immunology a source of joy and excitement, and to our families who made these careers possible. We hope that future generations of immunology students will find this subject as fascinating and rewarding as we have. About the Authors All four authors are active scholars and teachers who have been/are recipients of research grants from the NIH and the NSF. We have all served in various capacities as grant proposal reviewers for NSF, NIH, HHMI, and other funding bodies as well as evaluating manuscripts submitted for publication in immunological journals. In addition, we are all active members of the American Association of Immunologists and have served our national organization in a variety of ways. Judy Owen holds B.A. and M.A. (Hons) degrees from Cambridge University. She pursued her Ph.D. at the University of Pennsylvania with the late Dr. Norman Klinman and her postdoctoral fellowship with Dr. Peter Doherty in viral immunology. She was appointed to the faculty of Haverford College, one of the first undergraduate colleges to offer a course in immunology, in 1981. She teaches numerous laboratory and lecture courses in biochemistry and immunology and has received several teaching and mentorship awards. She is a participant in the First Year Writing Program and has been involved in curriculum development across the College. Jenni Punt received her A.B. from Bryn Mawr College (magna cum laude) majoring in Biology at Haverford College, She received her VMD (summa cum laude) and Ph.D. in immunology from the University of Pennsylvania and was a Damon Runyon-Walter Winchell Physician-Scientist fellow with Dr. Alfred Singer at the National Institutes of Health. She was appointed to the faculty of Haverford College in 1996 where she teaches cell biology and immunology and performs research in T cell development and hematopoiesis. She has received several teaching awards and has contributed to the development of college-wide curricular initiatives. Together, Jenni Punt and Judy Owen developed and ran the first AAI Introductory Immunology course, which is now offered on an annual basis. Sharon Stranford obtained her B.A. with Honors in Biology from Arcadia University and her Ph.D. in Microbiology and Immunology from Hahnemann (now Drexel) University, where she studied autoimmunity with funding from the Multiple Sclerosis Foundation. She pursued postdoctoral studies in transplantation immunology at Oxford University in England, followed by a fellowship at the University of California, San Francisco, working on HIV/AIDS with Dr. Jay Levy. From 1999 to 2001, Sharon was a Visiting Assistant Professor of Biology at Amherst College, and in 2001 joined the faculty of Mount Holyoke College as a Clare Boothe Luce Assistant Professor. She teaches courses in introductory biology, cell biology, immunology, and infectious disease, as well as a new interdisciplinary course called Controversies in Public Health. Pat Jones graduated from Oberlin College in Ohio with Highest Honors in Biology and obtained her Ph.D. in Biology with Distinction from the Johns Hopkins University. She was a postdoctoral fellow of the Arthritis Foundation for two years in the Department of Biochemistry and Biophysics at the University of California, San Francisco, Medical School, followed by two years as an NSF postdoctoral fellow in the Departments of Genetics and Medicine/ Immunology at Stanford University School of Medicine. In 1978 she was appointed Assistant Professor of Biology at Stanford and is now a full professor. Pat has received several undergraduate teaching awards, was the founding Director of the Ph.D. Program in Immunology, and in July, 2011, she assumed the position of Director of Stanford Immunology, a position that coordinates activities in immunology across the university. Contents Chapter 1 Overview of the Immune System A Historical Perspective of Immunology SUMMARY 23 REFERENCES 23 1 USEFUL WEB SITES 23 2 STUDY QUESTIONS 24 Early vaccination studies led the way to immunology 2 Vaccination is an ongoing, worldwide enterprise 3 Chapter 2 Immunology is about more than just vaccines and infectious disease 4 Immunity involves both humoral and cellular components 6 How are foreign substances recognized by the immune system? Cells, Organs, and Microenvironments of the Immune System 9 Important Concepts for Understanding the Mammalian Immune Response Pathogens come in many forms and must first breach natural barriers 11 12 The immune response quickly becomes tailored to suit the assault 12 Pathogen recognition molecules can be encoded in the germline or randomly generated 14 Tolerance ensures that the immune system avoids destroying the host 15 The immune response is composed of two interconnected arms: innate immunity and adaptive immunity 16 Adaptive immune responses typically generate memory 17 The Good, Bad, and Ugly of the Immune System 19 Cells of the Immune System 27 27 Hematopoietic stem cells have the ability to differentiate into many types of blood cells 28 Hematopoeisis is the process by which hematopoietic stem cells develop into mature blood cells 32 Cells of the myeloid lineage are the first responders to infection 32 Cells of the lymphoid lineage regulate the adaptive immune response 37 Primary Lymphoid Organs— Where Immune Cells Develop 41 The bone marrow provides niches for hematopoietic stem cells to self-renew and differentiate into myeloid cells and B lymphocytes 41 The thymus is a primary lymphoid organ where T cells mature 41 Secondary Lymphoid Organs— Where the Immune Response Is Initiated 48 Secondary lymphoid organs are distributed throughout the body and share some anatomical features 48 22 Lymphoid organs are connected to each other and to infected tissue by two different circulatory systems: blood and lymphatics 48 22 The lymph node is a highly specialized secondary lymphoid organ 50 Inappropriate or dysfunctional immune responses can result in a range of disorders 19 The immune response renders tissue transplantation challenging Cancer presents a unique challenge to the immune response viii Contents The spleen organizes the immune response against blood-borne pathogens 53 Signal-induced PIP2 breakdown by PLC causes an increase in cytoplasmic calcium ion concentration 75 MALT organizes the response to antigen that enters mucosal tissues 53 Ubiquitination may inhibit or enhance signal transduction 76 The skin is an innate immune barrier and also includes lymphoid tissue 56 Tertiary lymphoid tissues also organize and maintain an immune response 57 SUMMARY 60 REFERENCES Frequently Encountered Signaling Pathways 77 The PLC pathway induces calcium release and PKC activation 77 60 The Ras/Map kinase cascade activates transcription through AP-1 78 USEFUL WEB SITES 61 PKC activates the NF-κB transcription factor 79 STUDY QUESTIONS 61 Receptor-Ligand Interactions Receptor-ligand binding occurs via multiple noncovalent bonds 80 Antibodies share a common structure of two light chains and two heavy chains 81 There are two major classes of antibody light chains 85 There are five major classes of antibody heavy chains 85 66 Antibodies and antibody fragments can serve as antigens 86 66 Each of the domains of the antibody heavy and light chains mediate specific functions 88 X-ray crystallography has been used to define the structural basis of antigen-antibody binding 90 65 How do we quantitate the strength of receptorligand interactions? 66 Interactions between receptors and ligands can be multivalent 67 Receptor and ligand expression can vary during the course of an immune response Local concentrations of cytokines and other ligands may be extremely high Common Strategies Used in Many Signaling Pathways Ligand binding can induce conformational changes in, and/or clustering of, the receptor 80 Antibodies are made up of multiple immunoglobulin domains Chapter 3 Receptors and Signaling: B and T-Cell Receptors The Structure of Antibodies Signal Transduction in B Cells 91 68 Antigen binding results in docking of adapter molecules and enzymes into the BCR-Igα/Igβ membrane complex 91 68 B cells use many of the downstream signaling pathways described above 92 69 B cells also receive signals through co-receptors 94 T-Cell Receptors and Signaling 71 95 The T-cell receptor is a heterodimer with variable and constant regions 95 Some receptors require receptor-associated molecules to signal cell activation 71 The T-cell signal transduction complex includes CD3 98 Ligand-induced receptor clustering can alter receptor location 71 The T cell co-receptors CD4 and CD8 also bind the MHC 99 Tyrosine phosphorylation is an early step in many signaling pathways 73 Lck is the first tyrosine kinase activated in T cell signaling 100 Adapter proteins gather members of signaling pathways 74 T cells use downstream signaling strategies similar to those of B cells 100 Phosphorylation on serine and threonine residues is also a common step in signaling pathways SUMMARY 101 74 REFERENCES 102 USEFUL WEB SITES 102 STUDY QUESTIONS 103 Phosphorylation of membrane phospholipids recruits PH domain-containing proteins to the cell membrane 75 Contents Cytokine storms may have caused many deaths in the 1918 Spanish influenza Chapter 4 Receptors and Signaling: Cytokines and Chemokines General Properties of Cytokines and Chemokines Cytokine-Based Therapies 105 106 REFERENCES 138 USEFUL WEB SITES 139 STUDY QUESTIONS 140 Cytokines have numerous biological functions 107 Chapter 5 Cytokines can elicit and support the activation of specific T-cell subpopulations 107 Innate Immunity Cell activation may alter the expression of receptors and adhesion molecules 109 Signaling through multiple receptors can fine tune a cellular response Six Families of Cytokines and Associated Receptor Molecules Cytokines of the IL-1 family promote proinflammatory signals 110 113 Hematopoietin (Class I) family cytokines share three-dimensional structural motifs, but induce a diversity of functions in target cells 116 The Interferon (Class II) cytokine family was the first to be discovered 119 Members of the TNF cytokine family can signal development, activation, or death 123 The IL-17 family is a recently discovered, proinflammatory cytokine cluster 127 Chemokines direct the migration of leukocytes through the body 129 Cytokine Antagonists Anatomical Barriers to Infection 143 143 Antimicrobial proteins and peptides kill would-be invaders 145 147 Microbes are recognized by receptors on phagocytic cells 147 Phagocytosed microbes are killed by multiple mechanisms 151 Phagocytosis contributes to cell turnover and the clearance of dead cells 152 Induced Cellular Innate Responses 133 141 Epithelial barriers prevent pathogen entry into the body’s interior Phagocytosis 111 137 138 107 110 137 SUMMARY Cytokines mediate the activation, proliferation, and differentiation of target cells Cytokines are concentrated between secreting and target cells ix 152 Cellular pattern recognition receptors activate responses to microbes and cell damage 153 Toll-like receptors recognize many types of pathogen molecules 153 C-type lectin receptors bind carbohydrates on the surfaces of extracellular pathogens 158 Retinoic acid-inducible gene-I-like receptors bind viral RNA in the cytosol of infected cells 160 The IL-1 receptor antagonist blocks the IL-1 cytokine receptor 133 Nod-like receptors are activated by a variety of PAMPs, DAMPs, and other harmful substances 160 Cytokine antagonists can be derived from cleavage of the cytokine receptor 134 Expression of innate immunity proteins is induced by PRR signaling 160 Some viruses have developed strategies to exploit cytokine activity 134 Cytokine-Related Diseases Inflammatory Responses 134 166 Inflammation results from innate responses triggered by infection, tissue damage, or harmful substances 167 Proteins of the acute phase response contribute to innate immunity and inflammation 168 Septic shock is relatively common and potentially lethal 135 Bacterial toxic shock is caused by superantigen induction of T-cell cytokine secretion 135 Natural Killer Cells 168 Cytokine activity is implicated in lymphoid and myeloid cancers 137 Regulation and Evasion of Innate and Inflammatory Responses 169 x Contents Innate and inflammatory responses can be harmful 169 Innate and inflammatory responses are regulated both positively and negatively 172 Complement activity is passively regulated by protein stability and cell surface composition 210 Pathogens have evolved mechanisms to evade innate and inflammatory responses 173 The C1 inhibitor, C1INH, promotes dissociation of C1 components 211 Decay Accelerating Factors promote decay of C3 convertases 211 Interactions Between the Innate and Adaptive Immune Systems 173 The Regulation of Complement Activity 210 The innate immune system activates and regulates adaptive immune responses Factor I degrades C3b and C4b 212 174 Protectin inhibits the MAC attack 213 Adjuvants activate innate immune responses to increase the effectiveness of immunizations 175 Carboxypeptidases can inactivate the anaphylatoxins, C3a and C5a 213 Some pathogen clearance mechanisms are common to both innate and adaptive immune responses 176 Ubiquity of Innate Immunity 176 Complement Deficiencies 213 Microbial Complement Evasion Strategies 214 Plants rely on innate immune responses to combat infections 177 Some pathogens interfere with the first step of immunoglobulin-mediated complement activation 215 Invertebrate and vertebrate innate immune responses show both similarities and differences 177 Microbial proteins bind and inactivate complement proteins 215 SUMMARY 180 Microbial proteases destroy complement proteins 215 REFERENCES 181 USEFUL WEB SITES 182 Some microbes mimic or bind complement regulatory proteins 215 STUDY QUESTIONS 182 Chapter 6 The Complement System 187 The Evolutionary Origins of the Complement System 215 SUMMARY 219 REFERENCES 220 USEFUL WEB SITES 220 STUDY QUESTIONS 221 The Major Pathways of Complement Activation 189 The classical pathway is initiated by antibody binding 190 The lectin pathway is initiated when soluble proteins recognize microbial antigens 195 The alternative pathway is initiated in three distinct ways 196 The three complement pathways converge at the formation of the C5 convertase The Organization and Expression of Lymphocyte Receptor Genes 225 200 The Puzzle of Immunoglobulin Gene Structure C5 initiates the generation of the MAC 200 The Diverse Functions of Complement 201 Complement receptors connect complementtagged pathogens to effector cells 201 Complement enhances host defense against infection 204 Complement mediates the interface between innate and adaptive immunities 207 Complement aids in the contraction phase of the immune response 207 Complement mediates CNS synapse elimination 210 Chapter 7 226 Investigators proposed two early theoretical models of antibody genetics 226 Breakthrough experiments revealed that multiple gene segments encode the light chain 227 Multigene Organization of Ig Genes 231 Kappa light-chain genes include V, J, and C segments 231 Lambda light-chain genes pair each J segment with a particular C segment 231 Heavy-chain gene organization includes VH, D, JJ, and CH segments 232 Contents The Mechanism of V(D)J Recombination Recombination is directed by signal sequences Gene segments are joined by the RAG1/2 recombinase combination V(D)J recombination results in a functional Ig variable region gene V(D)J recombination can occur between segments transcribed in either the same or opposite directions Five mechanisms generate antibody diversity in naïve B cells B-Cell Receptor Expression Allelic exclusion ensures that each B cell synthesizes only one heavy chain and one light chain 232 233 234 235 239 239 242 242 Receptor editing of potentially autoreactive receptors occurs in light chains 243 Ig gene transcription is tightly regulated 244 Mature B cells express both IgM and IgD antibodies by a process that involves mRNA splicing 246 T-Cell Receptor Genes and Expression 247 Understanding the protein structure of the TCR was critical to the process of discovering the genes 247 The β-chain gene was discovered simultaneously in two different laboratories 249 A search for the α-chain gene led to the γ-chain gene instead 250 TCR genes undergo a process of rearrangement very similar to that of Ig genes 251 TCR expression is controlled by allelic exclusion 253 TCR gene expression is tightly regulated 253 SUMMARY 255 REFERENCES 256 USEFUL WEB SITES 257 STUDY QUESTIONS 258 The Major Histocompatibility Complex and Antigen Presentation The Structure and Function of MHC Molecules Class I molecules have a glycoprotein heavy chain and a small protein light chain Class II molecules have two non-identical glycoprotein chains 262 Class I and II molecules exhibit polymorphism in the region that binds to peptides 263 General Organization and Inheritance of the MHC 261 262 262 267 The MHC locus encodes three major classes of molecules 268 The exon/intron arrangement of class I and II genes reflects their domain structure 270 Allelic forms of MHC genes are inherited in linked groups called haplotypes 270 MHC molecules are codominantly expressed 271 Class I and class II molecules exhibit diversity at both the individual and species levels 273 MHC polymorphism has functional relevance 276 The Role of the MHC and Expression Patterns 277 MHC molecules present both intracellular and extracellular antigens 278 MHC class I expression is found throughout the body 278 Expression of MHC class II molecules is primarily restricted to antigen-presenting cells 279 MHC expression can change with changing conditions 279 T cells are restricted to recognizing peptides presented in the context of self-MHC alleles 281 Evidence suggests different antigen processing and presentation pathways 284 The Endogenous Pathway of Antigen Processing and Presentation 285 Peptides are generated by protease complexes called proteasomes 285 Peptides are transported from the cytosol to the RER 285 Chaperones aid peptide assembly with MHC class I molecules 286 The Exogenous Pathway of Antigen Processing and Presentation Chapter 8 xi 288 Peptides are generated from internalized molecules in endocytic vesicles 288 The invariant chain guides transport of class II MHC molecules to endocytic vesicles 289 Peptides assemble with class II MHC molecules by displacing CLIP 289 Cross-Presentation of Exogenous Antigens 291 xii Contents Dendritic cells appear to be the primary crosspresenting cell type 292 Mechanisms and Functions of Cross-Presentation 292 Presentation of Nonpeptide Antigens 293 SUMMARY 295 REFERENCES 295 USEFUL WEB SITES 296 STUDY QUESTIONS 296 Chapter 9 T-Cell Development Early Thymocyte Development 299 Apoptosis allows cells to die without triggering an inflammatory response 318 Different stimuli initiate apoptosis, but all activate caspases 318 Apoptosis of peripheral T cells is mediated by the extrinsic (Fas) pathway 320 TCR-mediated negative selection in the thymus induces the intrinsic (mitochondria-mediated) apoptotic pathway 321 Bcl-2 family members can inhibit or induce apoptosis 321 SUMMARY 324 REFERENCES 325 USEFUL WEB SITES 326 STUDY QUESTIONS 327 301 Thymocytes progress through four double-negative stages 301 Chapter 10 Thymocytes can express either TCRαβ or TCRγδ receptors 302 B-Cell Development DN thymocytes undergo β-selection, which results in proliferation and differentiation 303 Positive and Negative Selection The Site of Hematopoiesis 304 Thymocytes “learn” MHC restriction in the thymus 305 T cells undergo positive and negative selection 305 Positive selection ensures MHC restriction 307 Negative selection (central tolerance) ensures self-tolerance 310 The selection paradox: Why don’t we delete all cells we positively select? 312 An alternative model can explain the thymic selection paradox 313 Do positive and negative selection occur at the same stage of development, or in sequence? 314 Lineage Commitment Several models have been proposed to explain lineage commitment Double-positive thymocytes may commit to other types of lymphocytes Exit from the Thymus and Final Maturation 329 330 The site of B-cell generation changes during gestation 330 Hematopoiesis in the fetal liver differs from that in the adult bone marrow 332 B-Cell Development in the Bone Marrow The stages of hematopoiesis are defined by cellsurface markers, transcription-factor expression, and immunoglobulin gene rearrangements 332 334 The earliest steps in lymphocyte differentiation culminate in the generation of a common lymphoid progenitor 337 The later steps of B-cell development result in commitment to the B-cell phenotype 339 Immature B cells in the bone marrow are exquisitely sensitive to tolerance induction 344 Many, but not all, self-reactive B cells are deleted within the bone marrow 345 314 B cells exported from the bone marrow are still functionally immature 345 316 Mature, primary B-2 B cells migrate to the lymphoid follicles 349 314 316 Other Mechanisms That Maintain Self-Tolerance 316 The Development of B-1 and Marginal-Zone B Cells 351 TREG cells negatively regulate immune responses 317 B-1 B cells are derived from a separate developmental lineage 351 Peripheral mechanisms of tolerance also protect against autoreactive thymocytes 318 Marginal-zone cells share phenotypic and functional characteristics with B-1 B cells and arise at the T2 stage 352 Apoptosis 318 Comparison of B- and T-Cell Development 352 Contents SUMMARY 354 REFERENCES 355 USEFUL WEB SITES 355 STUDY QUESTIONS 356 Chapter 12 B-Cell Activation, Differentiation, and Memory Generation 385 T-Dependent B-Cell Responses Chapter 11 T-Cell Activation, Differentiation, and Memory 357 T-Cell Activation and the Two-Signal Hypothesis 358 Costimulatory signals are required for optimal T-cell activation and proliferation 359 Clonal anergy results if a costimulatory signal is absent 363 Cytokines provide Signal 3 364 Antigen-presenting cells have characteristic costimulatory properties 365 Superantigens are a special class of T-cell activators 366 T-Cell Differentiation 368 Helper T cells can be divided into distinct subsets 370 The differentiation of T helper cell subsets is regulated by polarizing cytokines 371 Effector T helper cell subsets are distinguished by three properties 372 Helper T cells may not be irrevocably committed to a lineage 378 Helper T-cell subsets play critical roles in immune health and disease 378 T-Cell Memory 379 xiii 388 T-dependent antigens require T-cell help to generate an antibody response 388 Antigen recognition by mature B cells provides a survival signal 389 B cells encounter antigen in the lymph nodes and spleen 390 B-cell recognition of cell-bound antigen results in membrane spreading 391 What causes the clustering of the B-cell receptors upon antigen binding? 392 Antigen receptor clustering induces internalization and antigen presentation by the B cell 393 Activated B cells migrate to find antigen-specific T cells 393 Activated B cells move either into the extrafollicular space or into the follicles to form germinal centers 395 Plasma cells form within the primary focus 395 Other activated B cells move into the follicles and initiate a germinal center response 396 Somatic hypermutation and affinity selection occur within the germinal center 398 Class switch recombination occurs within the germinal center after antigen contact 401 Most newly generated B cells are lost at the end of the primary immune response 403 Naïve, effector, and memory T cells display broad differences in surface protein expression 379 Some germinal center cells complete their maturation as plasma cells 403 TCM and TEM are distinguished by their locale and commitment to effector function 380 B-cell memory provides a rapid and strong response to secondary infection 404 How and when do memory cells arise? 380 What signals induce memory cell commitment? 381 Do memory cells reflect the heterogeneity of effector cells generated during a primary response? 381 Are there differences between CD4+ and CD8+ memory T cells? 381 How are memory cells maintained over many years? 381 SUMMARY 381 REFERENCES 382 USEFUL WEB SITES 383 STUDY QUESTIONS 383 T-Independent B-Cell Responses 406 T-independent antigens stimulate antibody production without the need for T-cell help 406 Two novel subclasses of B cells mediate the response to T-independent antigens 407 Negative Regulation of B Cells 411 Negative signaling through CD22 shuts down unnecessary BCR signaling 411 Negative signaling through the FcγRIIb receptor inhibits B-cell activation 411 B-10 B cells act as negative regulators by secreting IL-10 411 xiv Contents SUMMARY 412 REFERENCES 413 USEFUL WEB SITES 414 STUDY QUESTIONS 414 Naïve lymphocytes sample stromal cells in the lymph nodes 461 Naïve lymphocytes browse for antigen along reticular networks in the lymph node 461 Immune Cell Behavior during the Innate Immune Response 464 Chapter 13 Antigen-presenting cells travel to lymph nodes and present processed antigen to T cells 465 Effector Responses: Cell-and Antibody-Mediated Immunity Unprocessed antigen also gains access to lymphnode B cells 465 Antibody-Mediated Effector Functions Antibodies mediate the clearance and destruction of pathogen in a variety of ways Antibody isotypes mediate different effector functions Fc receptors mediate many effector functions of antibodies Cell-Mediated Effector Responses 415 416 Immune Cell Behavior during the Adaptive Immune Response 467 + 416 419 423 427 Naïve CD4 T cells arrest their movements after engaging antigens 468 + B cells seek help from CD4 T cells at the border between the follicle and paracortex of the Lymph Node 468 Dynamic imaging approaches have been used to address a controversy about B-cell behavior in germinal centers 470 + Cytotoxic T lymphocytes recognize and kill infected or tumor cells via T-cell receptor activation 428 CD8 T cells are activated in the lymph node via a multicellular interaction 471 Natural killer cells recognize and kill infected cells and tumor cells by their absence of MHC class I 435 Activated lymphocytes exit the lymph node and recirculate 472 A summary of our current understanding 472 The immune response contracts within 10 to 14 days 474 NKT cells bridge the innate and adaptive immune systems 441 Experimental Assessment of Cell-Mediated Cytotoxicity 444 Co-culturing T cells with foreign cells stimulates the mixed-lymphocyte reaction 444 Chemokine receptors and integrins regulate homing of effector lymphocytes to peripheral tissues 474 CTL activity can be demonstrated by cell-mediated lympholysis 445 Effector lymphocytes respond to antigen in multiple tissues 475 The graft-versus-host reaction is an in vivo indication of cell-mediated cytotoxicity 446 SUMMARY 480 REFERENCES 481 SUMMARY 446 USEFUL WEB SITES 482 REFERENCES 447 STUDY QUESTIONS 482 USEFUL WEB SITES 448 STUDY QUESTIONS 448 Immune Cell Behavior in Peripheral Tissues 474 Chapter 15 Chapter 14 The Immune Response in Space and Time Immune Cell Behavior before Antigen Is Introduced Naïve lymphocytes circulate between secondary and tertiary lymphoid tissues 451 455 455 Allergy, Hypersensitivities, and Chronic Inflammation 485 Allergy: A Type I Hypersensitivity Reaction 486 IgE antibodies are responsible for type I hypersensitivity 487 Many allergens can elicit a type I response 487 IgE antibodies act by cross-linking Fcε receptors on the surfaces of innate immune cells 487 Contents IgE receptor signaling is tightly regulated 491 Chapter 16 Innate immune cells produce molecules responsible for type I hypersensitivity symptoms 491 Type I hypersensitivities are characterized by both early and late responses 494 Tolerance, Autoimmunity, and Transplantation There are several categories of type I hypersensitivity reactions 494 There is a genetic basis for type I hypersensitivity 497 Diagnostic tests and treatments are available for type I hypersensitivity reactions The hygiene hypothesis has been advanced to explain increases in allergy incidence Antibody-Mediated (Type II) Hypersensitivity Reactions Transfusion reactions are an example of type II hypersensitivity 498 501 501 Hemolytic disease of the newborn is caused by type II reactions 503 Hemolytic anemia can be drug induced 504 Immune Complex-Mediated (Type III) Hypersensitivity Immune complexes can damage various tissues Immune complex-mediated hypersensitivity can resolve spontaneously 505 505 505 517 Establishment and Maintenance of Tolerance 518 Antigen sequestration is one means to protect self antigens from attack 519 Central tolerance limits development of autoreactive T cells and B cells 520 Peripheral tolerance regulates autoreactive cells in the circulation 520 Autoimmunity 501 xv 525 Some autoimmune diseases target specific organs 526 Some autoimmune diseases are systemic 529 Both intrinsic and extrinsic factors can favor susceptibility to autoimmune disease 531 Several possible mechanisms have been proposed for the induction of autoimmunity 533 Autoimmune diseases can be treated by general or pathway-specific immunosuppression 534 Transplantation Immunology 536 Graft rejection occurs based on immunologic principles 536 Graft rejection follows a predictable clinical course 541 Autoantigens can be involved in immune complexmediated reactions 506 Immunosuppressive therapy can be either general or target-specific 543 Arthus reactions are localized type III hypersensitivity reactions 506 Immune tolerance to allografts is favored in certain instances 545 Delayed-Type (Type IV) Hypersensitivity (DTH) 506 The initiation of a type IV DTH response involves sensitization by antigen Some organs are more amenable to clinical transplantation than others 546 507 SUMMARY 549 The effector phase of a classical DTH response is induced by second exposure to a sensitizing antigen 507 REFERENCES 550 USEFUL WEB SITES 551 The DTH reaction can be detected by a skin test 508 STUDY QUESTIONS 551 Contact dermatitis is a type IV hypersensitivity response 508 Chronic Inflammation Infections can cause chronic inflammation 509 509 Chapter 17 There are noninfectious causes of chronic inflammation 510 Infectious Diseases and Vaccines 553 Obesity is associated with chronic inflammation 510 Chronic inflammation can cause systemic disease 510 The Importance of Barriers to Infection and the Innate Response 554 SUMMARY 513 Viral Infections 555 REFERENCES 515 Many viruses are neutralized by antibodies 556 USEFUL WEB SITES 515 STUDY QUESTIONS 516 Cell-mediated immunity is important for viral control and clearance 556 xvi Contents Viruses employ several different strategies to evade host defense mechanisms 556 B-cell immunodeficiencies exhibit depressed production of one or more antibody isotypes 601 Influenza has been responsible for some of the worst pandemics in history 557 Disruptions to innate components may also impact adaptive responses 601 Complement deficiencies are relatively common 603 560 Immunodeficiency that disrupts immune regulation can manifest as autoimmunity 603 Bacteria can evade host defense mechanisms at several different stages 563 Immunodeficiency disorders are treated by replacement therapy 604 Tuberculosis is primarily controlled by CD4+ T cells 564 Animal models of immunodeficiency have been used to study basic immune function 604 Diphtheria can be controlled by immunization with inactivated toxoid 565 Bacterial Infections Immune responses to extracellular and intracellular bacteria can differ Parasitic Infections Protozoan parasites account for huge worldwide disease burdens A variety of diseases are caused by parasitic worms (helminths) Fungal Infections 560 565 565 567 569 Innate immunity controls most fungal infections 569 Immunity against fungal pathogens can be acquired 571 Emerging and Re-emerging Infectious Diseases 571 Some noteworthy new infectious diseases have appeared recently 572 Diseases may re-emerge for various reasons 573 Vaccines Protective immunity can be achieved by active or passive immunization 574 574 Secondary Immunodeficiencies 606 HIV/AIDS has claimed millions of lives worldwide 607 The retrovirus HIV-1 is the causative agent of AIDS 608 HIV-1 is spread by intimate contact with infected body fluids 610 In vitro studies have revealed the structure and life cycle of HIV-1 612 Infection with HIV-1 leads to gradual impairment of immune function 615 Active research investigates the mechanism of progression to AIDS 616 Therapeutic agents inhibit retrovirus replication 619 A vaccine may be the only way to stop the HIV/AIDS epidemic 621 SUMMARY 623 REFERENCES 623 USEFUL WEB SITES 624 STUDY QUESTIONS 624 There are several vaccine strategies, each with unique advantages and challenges 578 Conjugate or multivalent vaccines can improve immunogenicity and outcome 583 Chapter 19 Adjuvants are included to enhance the immune response to a vaccine 585 Cancer and the Immune System 627 SUMMARY 586 Terminology and Common Types of Cancer 627 REFERENCES 587 Malignant Transformation of Cells 628 USEFUL WEB SITES 588 DNA alterations can induce malignant transformation 629 STUDY QUESTIONS 588 The discovery of oncogenes paved the way for our understanding of cancer induction 629 Genes associated with cancer control cell proliferation and survival 630 Malignant transformation involves multiple steps 633 Chapter 18 Immunodeficiency Disorders Primary Immunodeficiencies Combined immunodeficiencies disrupt adaptive immunity 593 593 597 Tumor Antigens 634 Tumor-specific antigens are unique to tumor cells 636 Tumor-associated antigens are normal cellular proteins with unique expression patterns 636 Contents The Immune Response to Cancer Immunoediting both protects against and promotes tumor growth 638 639 Key immunologic pathways mediating tumor eradication have been identified 639 Some inflammatory responses can promote cancer 642 Some tumor cells evade immune recognition and activation 643 Cancer Immunotherapy 644 xvii Hemagglutination inhibition reactions are used to detect the presence of viruses and of antiviral antibodies 658 Bacterial agglutination can be used to detect antibodies to bacteria 659 Antibody Assays Based on Antigen Binding to Solid-Phase Supports 659 Radioimmunoassays are used to measure the concentrations of biologically relevant proteins and hormones in bodily fluids 659 Monoclonal antibodies can be targeted to tumor cells 644 ELISA assays use antibodies or antigens covalently bound to enzymes 660 Cytokines can be used to augment the immune response to tumors 646 The design of an ELISA assay must consider various methodological options 662 Tumor-specific T cells can be expanded and reintroduced into patients 647 ELISPOT assays measure molecules secreted by individual cells 663 New therapeutic vaccines may enhance the anti-tumor immune response 647 Western blotting can identify a specific protein in a complex protein mixture 664 Methods to Determine the Affinity of AntigenAntibody Interactions 664 Manipulation of costimulatory signals can improve cancer immunity 647 Combination cancer therapies are yielding surprising results 648 SUMMARY 649 REFERENCES 650 USEFUL WEB SITES 650 STUDY QUESTIONS 651 Chapter 20 Experimental Systems and Methods Antibody Generation Polyclonal antibodies are secreted by multiple clones of antigen-specific B cells A monoclonal antibody is the product of a single stimulated B cell Monoclonal antibodies can be modified for use in the laboratory or the clinic Immunoprecipitation- Based Techniques 653 654 654 654 655 656 Immunoprecipitation can be performed in solution 656 Immunoprecipitation of soluble antigens can be performed in gel matrices 656 Immunoprecipitation allows characterization of cell-bound molecules 657 Agglutination Reactions Hemagglutination reactions can be used to detect antigen conjugated to the surface of red blood cells 658 any 658 Equilibrium dialysis can be used to measure antibody affinity for antigen 665 Surface plasmon resonance is commonly used for measurements of antibody affinity 667 Microscopic Visualization of Cells and Subcellular Structures 668 Immunocytochemistry and immunohistochemistry use enzyme-conjugated antibodies to create images of fixed tissues 668 Immunoelectron microscopy uses gold beads to visualize antibody-bound antigens 669 Immunofluorescence-Based Imaging Techniques 669 Fluorescence can be used to visualize cells and molecules 669 Immunofl uorescence microscopy uses antibodies conjugated with fluorescent dyes 669 Confocal fluorescence microscopy provides threedimensional images of extraordinary clarity 670 Multiphoton fluorescence microscopy is a variation of confocal microscopy 670 Intravital imaging allows observation of immune responses in vivo 671 Flow Cytometry 672 Magnetic Activated Cell Sorting 677 Cell Cycle Analysis 678 Tritiated (3H) thymidine uptake was one of the first methods used to assess cell division 678 xviii Contents Colorimetric assays for cell division are rapid and eliminate the use of radioactive isotopes 678 Bromodeoxyuridine-based assays for cell division use antibodies to detect newly synthesized DNA 678 Propidium iodide enables analysis of the cell cycle status of cell populations Carboxyfluorescein succinimidyl ester can be used to follow cell division Assays of Cell Death 678 679 679 The 51Cr release assay was the first assay used to measure cell death 679 Fluorescently labeled annexin V measures phosphatidyl serine in the outer lipid envelope of apoptotic cells 680 684 Knock-in and knockout technologies replace an endogenous with a nonfunctional or engineered gene copy 685 The cre/lox system enables inducible gene deletion in selected tissues 687 SUMMARY 689 REFERENCES 690 USEFUL WEB SITES 690 STUDY QUESTIONS 691 Appendix I The TUNEL assay measures apoptotically generated DNA fragmentation 680 Caspase assays measure the activity of enzymes involved in apoptosis 681 Biochemical Approaches Used to Elucidate Signal Transduction Pathways 681 Biochemical inhibitors are often used to identify intermediates in signaling pathways 681 Many methods are used to identify proteins that interact with molecules of interest 682 Whole Animal Experimental Systems Transgenic animals carry genes that have been artificially introduced CD Antigens A-1 Appendix II Cytokines B-1 Appendix III 682 Animal research is subject to federal guidelines that protect nonhuman research subjects 682 Inbred strains can reduce experimental variation 683 Congenic resistant strains are used to study the effects of particular gene loci on immune responses 684 Adoptive transfer experiments allow in vivo examination of isolated cell populations 684 Chemokines and Chemokine Receptors Glossary Answers to Study Questions Index C-1 G-1 AN-1 I-1 Feature Boxes in Kuby 7e Clinical Focus Classic Experiment Box 1.1 Box 2.1 Box 2.3 Box 3.1 Box 3.3 Box 6.1 Box 7.1 Box 1.2 Box 1.3 Box 2.2 Box 3.2 Box 4.2 Box 4.4 Box 5.2 Box 6.2 Box 7.3 Box 8.2 Box 8.4 Box 9.2 Box 9.3 Box 10.1 Box 11.2 Box 11.4 Box 13.1 Box 15.2 Box 15.3 Box 16.1 Box 16.2 Box 16.4 Box 17.1 Box 18.1 Box 19.1 Vaccine Controversy: What’s Truth and What’s Myth? p. 5 Passive Antibodies and the Iditarod p. 8 The Hygiene Hypothesis p. 20 Stem Cells—Clinical Uses and Potential p. 42 Defects in the B-Cell Signaling Protein Btk Lead to X-Linked Agammaglobulinemia p. 93 Therapy with Interferons p. 120 Cytokines and Obesity p. 136 Genetic Defects in Components of Innate and Inflammatory Responses Associated with Disease p. 170 The Complement System as a Therapeutic Target p. 208 Some Immunodeficiencies Result from Impaired Receptor Gene Recombination p. 255 MHC Alleles and Susceptibility to Certain Diseases p. 277 Deficiencies in TAP Can Lead to Bare Lymphocyte Syndrome p. 287 How Do T Cells That Cause Type 1 Diabetes Escape Negative Selection? p. 311 Failure of Apoptosis Causes Defective Lymphocyte Homeostasis p. 322 B-Cell Development in the Aging Individual p. 333 Costimulatory Blockade p. 364 What a Disease Reveals about the Physiological Role of TH17 Cells p. 376 Monoclonal Antibodies in the Treatment of Cancer p. 420 The Genetics of Asthma and Allergy p. 498 Type 2 Diabetes, Obesity, and Inflammation p. 511 It Takes Guts to Be Tolerant p. 523 Why Are Women More Susceptible Than Men to Autoimmunity? Gender Differences in Autoimmune Disease p. 528 Is There a Clinical Future for Xenotransplantation? p. 548 The 1918 Pandemic Influenza Virus: Should It Publish or Perish? p. 557 Prevention of Infant HIV Infection by AntiRetroviral Treatment p. 610 A Vaccine to Prevent Cervical Cancer, and More p. 637 Evolution Box 2.4 Box 5.3 Box 7.2 Box 8.1 Variations on Anatomical Themes p. 57 Plant Innate Immune Responses p. 178 Evolution of Recombined Lymphocyte Receptors p. 240 The Sweet Smell of Diversity p. 275 Box 8.3 Box 9.1 Box 10.3 Box 11.1 Box 12.1 Box 13.2 Box 15.1 Box 16.3 Isolating Hematopoietic Stem Cells p. 29 The Discovery of a Thymus—and Two p. 46 The Elucidation of Antibody Structure p. 82 The Discovery of the T-Cell Receptor p. 96 The Discovery of Properdin p. 198 Hozumi and Tonegawa’s Experiment: DNA Recombination Occurs in immunoglobulin Genes in Somatic Cells p. 227 Demonstration of the Self-MHC Restriction of CD8⫹ T Cells p. 282 Insights about Thymic Selection from the First TCR Transgenic Mouse Have Stood the Test of Time p. 308 The Stages of B-Cell Development: Characterization of the Hardy Fractions p. 342 Discovery of the First Costimulatory Receptor: CD28 p. 362 Experimental Proof That Somatic Hypermutation and Antigen- Induced Selection Occurred Within the Germinal Centers p. 399 Rethinking Immunological Memory: NK Cells Join Lymphocytes as Memory-Capable Cells p. 442 The Discovery and Identification of IgE as the Carrier of Allergic Hypersensitivity p. 488 Early Life Exposure to Antigens Favors Tolerance Induction p. 546 Advances Box 4.1 Box 4.3 Box 5.1 Box 6.3 Box 10.2 Box 11.3 Box 12.2 Box 14.1 Box 14.2 Box 17.2 Box 20.1 Methods Used to Map the Secretome p. 111 How Does Chemokine Binding to a Cell-Surface Receptor Result in Cellular Movement Along the Chemokine Gradient? p. 130 Inflammasomes p. 162 Staphylococcus aureus Employs Diverse Methods to Evade Destruction by the Complement System p. 216 The Role of miRNAs in the Control of B-Cell Development p. 336 How Many TCR Complexes Must Be Engaged to Trigger T-Cell Activation? p. 368 New Ideas on B-Cell Help: Not All Cells That Help B Cells Make Antibodies Are T Cells p. 408 Dynamic Imaging Techniques p. 452 Molecular Regulation of Cell Migration Between and Within Tissues p. 456 A Prime and Pull Vaccine Strategy for Preventing Sexually Transmitted Diseases p. 586 Flow Cytometry Under the Hood p. 674 Preface Like all of the previous authors of this book, we are dedicated to the concept that immunology is best taught and learned in an experimentally-based manner, and we have retained that emphasis with this edition. It is our goal that students should complete an immunology course not only with a firm grasp of content, but also with a clear sense of how key discoveries were made, what interesting questions remain, and how they might best be answered. We believe that this approach ensures that students both master fundamental immunological concepts and internalize a vision of immunology as an active and ongoing process. Guided by this vision, the new edition has been extensively updated to reflect the recent advances in all aspects of our discipline. New Authorship As a brand-new team of authors, we bring experience in both research and undergraduate teaching to the development of this new edition, which continues to reflect a dedication to pedagogical excellence originally modeled by Janis Kuby. We remain deeply respectful of Kuby’s unique contribution to the teaching of immunology and hope and trust that this new manifestation of her creation will simply add to her considerable legacy. 2a 1 Lymph node P 3 P P B B T T 5a T B 4 5b Memory 2b N B T OVERVIEW FIGURE 1-9 Collaboration between innate and adaptive immunity in resolving an infection. A new capstone chapter (Chapter 14) integrates the events of an immune response into a complete story, with particular reference to the advanced imaging techniques that have become available since the writing of the previous edition. In this way, the molecular and cellular details presented in Chapters 2-13 are portrayed in context, a moving landscape of immune response events in time and space (Figure 14-5). Understanding Immunology As a Whole We recognize that the immune system is an integrated network of cells, molecules, and organs, and that each component relies on the rest to function properly. This presents a pedagogical challenge because to understand the whole, we must attain working knowledge of many related pieces of information, and these do not always build upon each other in simple linear fashion. In acknowledgment of this challenge, this edition presents the “big picture” twice; first as an introductory overview to immunity, then, thirteen chapters later, as an integration of the details students have learned in the intervening text. Specifically, Chapter 1 has been revised to make it more approachable for students who are new to immunology. The chapter provides a short historical background to the field and an introduction to some of the key players and their roles in the immune response, keeping an eye on fundamental concepts (Overview Figure 1-9). A new section directly addresses some of the biggest conceptual hurdles, but leaves the cellular and molecular details for later chapters. FIGURE 14-5 A T cell (blue) on a fibroblastic reticular network (red and green) in the lymph node. xxi Preface Focus on the Fundamentals The order of chapters in the seventh edition has been revised to better reflect the sequence of events that occurs naturally during an immune response in vivo. This offers instructors the opportunity to lead their students through the steps of an immune response in a logical sequence, once they have learned the essential features of the tissues, cells, molecular structures, ligand-receptor binding interactions, and signaling pathways necessary for the functioning of the immune system. The placement of innate immunity at the forefront of the immune response enables it to take its rightful place as the first, and often the only, aspect of immunity that an organism needs to counter an immune insult. Similarly, the chapter on complement is located within the sequence in a place that highlights its function as a bridge between innate and adaptive immune processes. However, we recognize that a course in immunology is approached differently by each instructor. Therefore, as much as possible, we have designed each of the chapters so that it can stand alone and be offered in an alternative order. signaling, as well as to specific molecules and pathways involved in signaling through antigen receptors. Chapter 4 includes a more thorough introduction to the roles of cytokines and chemokines in the immune response. • An expanded and updated treatment of innate immunity (Chapter 5), which now includes comprehensive coverage of the many physical, chemical, and cellular defenses that constitute the innate immune system, as well as the ways in which it activates and regulates adaptive immunity. • Substantial rewriting of chapters concerned with complement (Chapter 6) and antigen receptor gene rearrangement (Chapter 7). These chapters have been extensively revised for clarity in both text and figures. The description of the complement system has been updated to include the involvement of complement proteins in both innate and adaptive aspects of immunity. • A restructured presentation of the MHC, with the addition of new information relevant to cross-presentation pathways (Chapter 8) (Figure 8-22b). (b) DC cross-presentation and activation of CTL Challenging All Levels While this book is written as a text for students new to immunology, it is also our intent to challenge students to reach deeply into the field and to appreciate the connections with other aspects of biology. Instead of reducing difficult topics to vague and simplistic forms, we instead present them with the level of detail and clarity necessary to allow the beginning student to find and understand information they may need in the future. This offers the upper level student a foundation from which they can progress to the investigation of advances and controversies within the current immunological literature. Supplementary focus boxes have been used to add nuance or detail to discussions of particular experiments or ideas without detracting from the flow of information. These boxes, which address experimental approaches, evolutionary connections, clinical aspects, or advanced material, also allow instructors to tailor their use appropriately for individual courses. They provide excellent launching points for more intensive in class discussions relevant to the material. Some of the most visible changes and improvements include: • A rewritten chapter on the cells and organs of the immune system (Chapter 2) that includes up to date images reflecting our new understanding of the microenvironments where the host immune system develops and responds. • The consolidation of signaling pathways into two chapters: Chapter 3 includes a basic introduction to ligand:receptor interactions and principles of receptor Cross-presenting dendritic cell Exogenous antigen TLR Crossover pathway Class I MHC CD8 CD3 CD80/CD86 CD28 IL-2 Naïve TC cell FIGURE 8-22b Exogenous antigen activation of naïve Tc cells requires DC licensing and cross-presentation • The dedication of specialized chapters concerned with T cell development and T cell activation (Chapters 9 and Chapter 11, respectively). Chapter 11 now includes current descriptions of the multiple helper T cell subsets that regulate the adaptive immune response. • Substantially rewritten chapters on B cell development and B cell activation (Chapters 10 and 12, respectively) that address the physiological locations as well as the nature of the interacting cells implicated in these processes. • An updated discussion of the role of effector cells and molecules in clearing infection (Chapter 13), including a more thorough treatment of NK and NKT cells. xxii Preface • A new chapter that describes advances in understanding and visualizing the dynamic behavior and activities of immune cells in secondary and tertiary tissue (Chapter 14). • Substantial revision and updating of the clinical chapters (Chapters 15-19) including the addition of several new clinically relevant focus boxes. • Revised and updated versions of the final methods chapter (Chapter 20), and the appendices of CD antigens, chemokines, and cytokines and their receptors. Throughout the book, we attempt to provide a “big picture” context for necessary details in a way that facilitates greater student understanding. Recent Advances and Other Additions Immunology is a rapidly growing field, with new discoveries, advances in techniques, and previously unappreciated connections coming to light every day. The 7th edition has been thoroughly updated throughout, and now integrates the following new material and concepts: • New immune cell types and subtypes, as well as the phenotypic plasticity that is possible between certain subtypes of immune cells. • A greater appreciation for the wide range of mechanisms responsible for innate immunity and the nature and roles of innate responses in sensing danger, inducing inflammation, and shaping the adaptive response (Figure 5-18). 2 T cell Inhibitory cytokines T cell TGF␤ 1 Cytokine deprivation IL–2R FoxP3 3 TCR APC Inhibiting antigen presenting cells MHC 4 Cytotoxicity T cell T cell FIGURE 9-10 How regulatory T cells inactivate traditional T cells. • The roles of the microbiome and commensal organisms in the development and function of immunity, as well as the connections between these and many chronic diseases. • A new appreciation for the micro environmental substructures that guide immune cell interactions with antigen and with one another (Figure 14-11a). Antigen delivery to T cells Subcapsular sinus (SCS) Lymph node DC presenting antigen Afferent lymphatic Antigen B cell follicle Bacteria Naïve TLR4 or TLR5 Dectin-1 TH1 IFN-γ IL-12 TLR3, 7, 9 Fungi IL-6 IL-23 Virus Naïve TH17 IL-17 Tricellular complex (CD8+ T cell, CD4+ T cell, and DC) FRC network T cell zone (paracortex) FIGURE 14-11a How antigen travels into a lymph node. IL-10 TLR2/1 Naïve Helminth TH2 IL-4 IL-5 IL-13 TLR2/6 Fungi IL-10 RA TGF-β CD28 CD80/86 TCR MHC II with peptide Naïve Treg IL-10 TGF-β FIGURE 5-18 Differential signaling through dendritic cell PRRs influences helper T cell functions. • Regulation of immunity, including new regulatory cell types, immunosuppressive chemical messengers and the roles these play, for example, in tolerance and in the nature of responses to different types of antigens (Figure 9-10). • Many technical advances, especially in the areas of imaging and sequencing, which have collectively enhanced our understanding of immune function and cellular interactions, allowing us to view the immune response in its natural anatomical context, and in real time (see Figure 14-5). Connections to the Bench, the Clinic, and Beyond We have made a concerted effort in the 7th edition to integrate experimental and clinical aspects of immunology into the text. In Chapter 2, illustrations of immune cells and tissues are shown alongside histological sections or, where possible, electron Preface micrographs, so students can see what they actually look like. Throughout the text, experimental data are used to demonstrate the bases for our knowledge (Figure 3-4b), and the clinical chapters at the end of the book (Chapters 15 through 19) describe new advances, new challenges, and newly appreciated connections between the immune system and disease. FIGURE 3-4b Targeted delivery of cytokines (pink). Featured Boxes Associated with each chapter are additional boxed materials that provide specialized information on historically-important studies (Classic Experiments) that changed the way immunologists viewed the field, noteworthy new breakthroughs (Advances) that have occurred since the last edition, the clinical relevance of particular topics (Clinical Focus) and the evolution of aspects of immune functioning (Evolution). Examples of such boxes are “The Prime and Pull Vaccine strategy,” “Genetic defects in components of innate and inflammatory responses associated with disease,” “The role of miRNAs in the control of B cell development” and an updated “Stem cells: Clinical uses and potential.” We have involved our own undergraduate students in the creation of some of these boxes, which we believe have greatly benefitted from their perspective on how to present interesting material effectively to their fellow students. Critical Thinking and Data Analysis Integration of experimental evidence throughout the book keeps students focused on the how and why. Detailed and clear descriptions of the current state of the field provide students with the knowledge, skills, and vocabulary to read critically in the primary literature. Updated and revised study questions at the end of the chapter range from simple recall of information to analyzing original data or proposing hypotheses to explain remaining questions in the field. Classic Experiment boxes throughout the text help students to appreciate the seminal experiments in immunology and how they were conducted, providing a bridge to the primary research articles and emphasizing data analysis at every step. Media and Supplements NEW! ImmunoPortal (courses.bfwpub.com/immunology7e) This comprehensive and robust online teaching and learning tool combines a wealth of media resources, vigorous xxiii assessment, and helpful course management features into one convenient, fully customizable space. ImmunoPortal Features: NEW! Kuby Immunology Seventh Edition e-Book—also available as a standalone resource (ebooks.bfwpub.com/ immunology7e) This online version of the textbook combines the contents of the printed book, electronic study tools, and a full complement of student media, including animations and videos. Students can personalize their e-Book with highlighting, bookmarking, and note-taking features. Instructors can customize the e-Book to focus on specific sections, and add their own notes and files to share with their class. NEW! LearningCurve—A Formative Quizzing Engine With powerful adaptive quizzing, a game-like format, and the promise of significantly better grades, LearningCurve gives instructors a quickly implemented, highly effective new way to get students more deeply involved in the classroom. Developed by experienced teachers and experts in educational technology, LearningCurve offers a series of brief, engaging activities specific to your course. These activities put the concept of “testing to learn” into action with adaptive quizzing that treats each student as an individual with specific needs: • Students work through LearningCurve activities one question at a time. • With each question, students get immediate feedback. Responses to incorrect answers include links to book sections and other resources to help students focus on what they need to learn. • As they proceed toward completion of the activity, the level of questioning adapts to the level of performance. The questions become easier, harder, or the same depending on how the student is doing. • And with a more confident understanding of assigned material, students will be more actively engaged during classtime. Resources The Resources center provides quick access to all instructor and student resources for Kuby Immunology. For Instructors— All instructor media are available in the ImmunoPortal and on the Instructor Resource DVD. NEW! test bank—over 500 dynamic questions in PDF and editable Word formats include multiple-choice and shortanswer problems, rated by level of difficulty and Bloom’s Taxonomy level. xxiv Acknowledgements Fully optimized JPEG files of every figure, photo, and table in the text, featuring enhanced color, higher resolution, and enlarged fonts. Images are also offered in PowerPoint® format for each chapter. Animations of complex text concepts and figures help students better understand key immunological processes. Videos specially chosen by the authors to complement and supplement text concepts. For Students— All of these resources are also available in the ImmunoPortal. • Student versions of the Animations and Videos, to help students understand key mechanisms and techniques at their own pace. • Flashcards test student mastery of vocabulary and allow students to tag the terms they’ve already learned. • Immunology on the Web weblinks introduce students to a world of online immunology resources and references. Assignments In this convenient space, ImmunoPortal provides instructors with the ability to assign any resource, as well as e-Book readings, discussion board posts, and their own materials. A gradebook tracks all student scores and can be easily exported to Excel or a campus Course Management System. Acknowledgements We owe special thanks to individuals who offered insightful ideas, who provided detailed reviews that led to major improvements, and who provided the support that made writing this text possible. These notable contributors include Dr. Stephen Emerson, Dr. David Allman, Dr. Susan Saidman, Dr. Nan Wang, Nicole Cunningham, and the many undergraduates who provided invaluable students’ perspectives on our chapters. We hope that the final product reflects the high quality of the input from these experts and colleagues and from all those listed below who provided critical analysis and guidance. We are also grateful to the previous authors of Kuby’s Immunology, whose valiant efforts we now appreciate even more deeply. Their commitment to clarity, to providing the most current material in a fast moving discipline, and to maintaining the experimental focus of the discussions set the standard that is the basis for the best of this text. We also acknowledge that this book represents the work not only of its authors and editors, but also of all those whose experiments and writing provided us with ideas, inspiration and information. We thank you and stress that all errors and inconsistencies of interpretation are ours alone. We thank the following reviewers for their comments and suggestions about the manuscript during preparation of this seventh edition. Their expertise and insights have contributed greatly to the book. Lawrence R. Aaronson, Utica College Jeffrey K. Actor, University of Texas Medical School at Houston Richard Adler, University of Michigan-Dearborn Emily Agard, York University, North York Karthik Aghoram, Meredith College Rita Wearren Alisauskas, Rutgers University John Allsteadt, Virginia Intermont College Gaylene Altman, University of Washington Angelika Antoni, Kutztown University Jorge N. Artaza, Charles R. Drew University of Medicine and Science Patricia S. Astry, SUNY Fredonia Roberta Attanasio, Georgia State University Elizabeth Auger, Saint Joseph’s College of Maine Avery August, Penn State University Rajeev Aurora, Saint Louis University Hospital Christine A. Bacon, Bay Path College Jason C. Baker, Missouri Western State College Kenneth Balazovich, University of Michigan-Dearborn Jennifer L. Bankers-Fulbright, Augsburg College Amorette Barber, Longwood University Brianne Barker, Hamilton College Scott R. Barnum, University of Alabama at Birmingham Laura Baugh, University of Dallas Marlee B. Marsh, Columbia College Rachel Venn Beecham, Mississippi Valley State University Fabian Benencia, Ohio University Main Campus Charlie Garnett Benson, Georgia State University Daniel Bergey, Black Hills State University Carolyn A. Bergman, Georgian Court College Elke Bergmann-Leitner, WRAIR/Uniformed Services University of Health Services Acknowledgements Brian P. Bergstrom, Muskingum College Susan Bjerke, Washburn University of Topeka Earl F. Bloch, Howard University Elliott J. Blumenthal, Indiana University–Purdue University Fort Wayne Kathleen Bode, Flint Hills Technical College Dennis Bogyo, Valdosta State University Mark Bolyard, Union University Lisa Borghesi, University of Pittsburgh Phyllis C. Braun, Fairfield University Jay H. Bream, Johns Hopkins University School of Medicine Heather A. Bruns, Ball State University Walter J. Bruyninckx, Hanover College Eric L. Buckles, Dillard University Sandra H. Burnett, Brigham Young University Peter Burrows, University of Alabama at Birmingham Ralph Butkowski, Augsburg College Jean A. Cardinale, Alfred University Edward A. Chaperon, Creighton University Stephen K. Chapes, Kansas State University Christopher Chase, South Dakota State University Thomas Chiles, Boston College Harold Chittum, Pikeville College Peter A. Chung, Pittsburg State University Felicia L. Cianciarulo, Carlow University Bret A. Clark, Newberry College Patricia A. Compagnone-Post, Albertus Magnus College Yasemin Kaya Congleton, Bluegrass Community and Technical College Vincent A. Connors, University of South CarolinaSpartanburg Conway-Klaassen, University of Minnesota Lisa Cuchara, Quinnipiac University Tanya R. Da Sylva, York University, North York Kelley L. Davis, Nova Southeastern University Jeffrey Dawson, Duke University Joseph DeMasi, Massachusetts College of Pharmacy & Allied Health Stephanie E. Dew, Centre College Joyce E. S. Doan, Bethel University Diane Dorsett, Georgia Gwinnett College James R. Drake, Albany Medical College Erastus C. Dudley, Huntingdon College Jeannine M. Durdik, University of Arkansas Fayetteville Karen M. Duus, Albany Medical College Christina K. Eddy, North Greenville University Anthony Ejiofor, Tennessee State University Jennifer Ellington, Belmont Abbey College Samantha L. Elliott, Saint Mary’s College of Maryland Lehman L. Ellis, Our Lady of Holy Cross College Sherine F. Elsawa, Northern Illinois University xxv Uthayashanker Ezekiel, Saint Louis University Medical Center Diana L. Fagan, Youngstown State University Rebecca V. Ferrell, Metropolitan State College of Denver Ken Field, Bucknell University Krista Fischer-Stenger, University of Richmond Howard B. Fleit, SUNY at Stony Brook Sherry D. Fleming, Kansas State University Marie-dominique Franco, Regis University Joel Gaikwad, Oral Roberts University D. L. Gibson, University of British Columbia-Okanagan Laura Glasscock, Winthrop University David Glick, Kings College Elizabeth Godrick, Boston University Karen Golemboski, Bellarmine University Sandra O. Gollnick, SUNY Buffalo James F. Graves, University of Detroit-Mercy Demetrius Peter Gravis, Beloit College Anjali D. Gray, Lourdes University Valery Z. Grdzelishvili, University of North Carolina-Charlotte Carla Guthridge, Cameron University David J. Hall, Lawrence University Sandra K. Halonen, Montana State University Michael C. Hanna, Texas A & M-Commerce Kristian M. Hargadon, Hampden-Sydney College JL Henriksen, Bellevue University Michelle L. Herdman, University of Charleston Jennifer L. Hess, Aquinas College Edward M. Hoffmann, University of Florida Kristin Hogquist, University of Minnesota Jane E. Huffman, East Stroudsburg University of Pennsylvania Lisa A. Humphries, University of California, Los Angeles Judith Humphries, Lawrence University Mo Hunsen, Kenyon College Vijaya Iragavarapu-Charyulu, Florida Atlantic University Vida R. Irani, Indiana University of Pennsylvania Christopher D. Jarvis, Hampshire College Eleanor Jator, Austin Peay State University Stephen R. Jennings, Drexel University College of Medicine Robert Jonas, Texas Lutheran University Vandana Kalia, Penn State University- Main Campus Azad K. Kaushik, University of Guelph George Keller, Samford University Kevin S. Kinney, De Pauw University Edward C. Kisailus, Canisius College David J. Kittlesen, University of Virginia Dennis J. Kitz, Southern Illinois University-Edwardsville Janet Kluftinger, University of British Columbia -Okanagan Rolf König, University of Texas Medical Branch at Galveston Kristine Krafts, University of Minnesota-Duluth Ruhul Kuddus, Utah Valley University Narendra Kumar, Texas A&M Health Science Center xxvi Acknowledgements N. M. Kumbaraci, Stevens Institute of Technology Jesse J. Kwiek, The Ohio State University Main Camp John M. Lammert, Gustavus Aldolphus College Courtney Lappas, Lebanon Valley College Christopher S. Lassiter, Roanoke College Jennifer Kraft Leavey, Georgia Institute of Technology Melanie J. Lee-Brown, Guilford College Vicky M. Lentz, SUNY College at Oneonta Joseph Lin, Sonoma State University Joshua Loomis, Nova Southeastern University Jennifer Louten, Southern Polytechnic State University Jon H. Lowrance, Lipscomb University Milson J. Luce, West Virginia University Institute of Technology Phillip J. Lucido, Northwest Missouri State University M.E. MacKay, Thompson Rivers University Andrew P. Makrigiannis, University of Ottawa Greg Maniero, Stonehill College David Markwardt, Ohio Wesleyan University John Martinko, Southern Illinois University Andrea M. Mastro, Penn State University-Main Campus Ann H. McDonald, Concordia University Lisa N. McKernan, Chestnut Hill College Catherine S. McVay, Auburn University Daniel Meer, Cardinal Stritch University JoAnn Meerschaert, Saint Cloud State University Brian J. Merkel, University Wisconsin-Green Bay Jiri Mestecky, University of Alabama at Birmingham Dennis W. Metzger, Albany Medical College Jennifer A. Metzler, Ball State University John A. Meyers, Boston University Medical School Yuko J. Miyamoto, Elon College Jody M. Modarelli, Hiram College Devonna Sue Morra, Saint Francis University Rita B. Moyes, Texas A&M Annette Muckerheide, College of Mount Saint Joseph Sue Mungre, Northeastern Illinois University Kari L. Murad, College of Saint Rose Karen Grandel Nakaoka, Weber State University Rajkumar Nathaniel, Nicholls State University David Nemazee, University of California, San Diego Hamida Rahim Nusrat, San Francisco State University Tracy O’Connor, Mount Royal College Marcos Oliveira, University of the Incarnate Word Donald Ourth, University of Memphis Deborah Palliser, Albert Einstein College of Medicine Shawn Phippen, Valdosta State University Melinda J. Pomeroy-Black, La Grange College Edith Porter, California State University, Los Angeles Michael F. Princiotta, SUNY Upstate Medical University Gerry A Prody, Western Washington University Robyn A. Puffenbarger, Bridgewater College Aimee Pugh-Bernard, University of Colorado at Denver Pattle Pun, Wheaton College Sheila Reilly, Belmont Abbey College Karen A. Reiner, Andrews University Margaret Reinhart, University of the Sciences in Philadelphia Stephanie Richards, Bates College Sarah M. Richart, Azusa Pacific University James E. Riggs, Rider University Vanessa Rivera-Amill, Ponce School of Medicine Katherine Robertson, Westminster College James L. Rooney, Lincoln University Robin S. Salter, Oberlin College Sophia Sarafova, Davidson College Surojit Sarkar, Penn State University-Main Campus Perry M. Scanlan, Austin Peay State University Ralph Seelke, University Wisconsin-Superior Diane L. Sewell, University Wisconsin-La Crosse Anding Shen, Calvin College Penny Shockett, Southeastern Louisiana University Michael Sikes, North Carolina State University Maryanne C. Simurda, Washington and Lee University Paul K. Small, Eureka College Jonathan Snow, Williams College Ralph A. Sorensen, Gettysburg College Andrew W. Stadnyk, Dalhousie University Faculty of Medicine Douglas A. Steeber, University Wisconsin-Milwaukee Viktor Steimle, University of Sherbrooke, Sherbrooke Douglas J. Stemke, University of Indianapolis Carolyn R. Stenbak, Seattle University Jennifer Ripley Stueckle, West Virginia University Kathleen Sullivan, Louisiana Technical College Alexandria Susmit Suvas, Oakland University Gabor Szalai, University of South Carolina Seetha M Tamma, Long Island University-C.W. Post Matthew J. Temple, Nazareth College Kent R. Thomas, Wichita State University Diane G. Tice, SUNY Morrisville Sara Sybesma Tolsma, Northwestern College Clara Tóth, Saint Thomas Aquinas College Bebhinn Treanor, University of Toronto Scarborough Allen W. Tsang, Bowman Gray Medical School Amar S. Tung, Lincoln University Lloyd Turtinen, University Wisconsin-Eau Claire Timothy M VanWagoner, Oklahoma Christian/ University of Oklahoma HSC Evros Vassiliou, Kean University Vishwanath Venketaraman, Western University of Health Sciences Kathleen Verville, Washington College Katherine A. Wall, University of Toledo Helen Walter, Mills College Acknowledgements Christopher Ward, University of Alberta Benjamin S. Weeks, Adelphi University Ben B. Whitlock, University of Saint Francis Robert Winn, Northern Michigan University Candace R. Winstead, California Polytechnic State UniversitySan Luis Obispo Dorothy M. Wrigley, Minnesota State University Jodi L. Yorty, Elizabethtown College Sheryl Zajdowicz, Metropolitan State University of Denver Mary Katherine Zanin, The Citadel The Military College of South Carolina Gary Zieve, SUNY at Stony Brook Michael I. Zimmer, Purdue Calumet Gilbert L. Zink, University of the Sciences in Philadelphia Patty Zwollo, College of William & Mary xxvii Finally, we thank our experienced and talented colleagues at W. H. Freeman and Company. Particular thanks to the production team members Philip McCaffrey, Sherrill Redd, Heath Lynn Silberfeld, Diana Blume, Lawrence Guerra, Janice Donnola, Christine Buese, and Elyse Reider. Thanks are also due to the editorial team of Lauren Schultz, Susan Winslow, Allison Michael, Yassamine Ebadat, and Irene Pech. However, a very special thanks go to our developmental editor, Erica Champion, and our developmental coordinator, Sara Ruth Blake. Erica has guided us from the beginning with a probing vision, endless patience, and keen eye for narrative and clarity. Sara kept us organized and true to deadlines with heroic resolve. The involvement of these two extraordinarily talented team members has made this edition, and its ambitious aspirations, possible. This page lelt intentionally blank. 1 Overview of the Immune System T he immune system evolved to protect multicellular organisms from pathogens. Highly adaptable, it defends the body against invaders as diverse as the tiny (~30 nm), intracellular virus that causes polio and as large as the giant parasitic kidney worm Dioctophyme renale, which can grow to over 100 cm in length and 10 mm in width. This diversity of potential pathogens requires a range of recognition and destruction mechanisms to match the multitude of invaders. To accomplish this feat, vertebrates have evolved a complicated and dynamic network of cells, molecules, and pathways. Although elements of these networks can be found throughout the plant and animal kingdoms, the focus of this book will be on the highly evolved mammalian immune system. The fully functional immune system involves so many organs, molecules, cells, and pathways in such an interconnected and sometimes circular process that it is often difficult to know where to start! Recent advances in cell imaging, genetics, bioinformatics, as well as cell and molecular biology, have helped us to understand many of the individual players in great molecular detail. However, a focus on the details (and there are many) can make taking a step back to see the bigger picture challenging, and it is often the bigger picture that motivates us to study immunology. Indeed, the field of immunology can be credited with the vaccine that eradicated smallpox, the ability to transplant organs between humans, and the drugs used today to treat asthma. Our goal in this chapter is therefore to present the background and concepts in immunology that will help bridge the gap between the cellular and molecular detail presented in subsequent chapters and the complete picture of an immune response. A clear understanding of each of the many players involved will help one appreciate the intricate coordination of an immune system that makes all of this possible. The study of immunology has produced amazing and fascinating stories (some of which you will see in this book), where host and microbe engage in battles waged over both minutes and millennia. But the immune system is also much more than an isolated component of the body, merely responsible for search-and-destroy missions. In fact, it interleaves with many of the other body systems, A phagocytic cell (macrophage, green) engulfing the bacteria that cause tuberculosis (orange). Max Planck Institute for Infection Biology/Dr. Volker Brinkmann ■ ■ ■ A Historical Perspective of Immunology Important Concepts for Understanding the Mammalian Immune Response The Good, Bad, and Ugly of the Immune System including the endocrine, nervous, and metabolic systems, with more connections undoubtedly to be discovered in time. Finally, it has become increasingly clear that elements of immunity play key roles in regulating homeostasis in the body for a healthy balance. Information gleaned from the study of the immune system, as well as its connections with other systems, will likely have resounding repercussions across many basic science and biomedical fields, not to mention in the future of clinical medicine. This chapter begins with a historical perspective, charting the beginnings of the study of immunology, largely driven by the human desire to survive major outbreaks of infectious disease. This is followed by presentation of a few key concepts that are important hallmarks of the mammalian immune response, many of which may not have been encountered elsewhere in basic biology. This is not meant as a comprehensive overview of the mammalian immune system but rather as a means for jumping the large conceptual hurdles frequently encountered as one begins to describe the complexity and interconnected nature of the immune response. We hope this will whet the appetite and prepare the reader for a more thorough discussion of the specific components of immunity presented in the 1 2 PA R T I | Introduction following chapters. We conclude with a few challenging clinical situations, such as instances in which the immune system fails to act or becomes the aggressor, turning its awesome powers against the host. More in-depth coverage of these and other medical aspects of immunology can be found in the final chapters of this book. A Historical Perspective of Immunology The discipline of immunology grew out of the observation that individuals who had recovered from certain infectious diseases were thereafter protected from the disease. The Latin term immunis, meaning “exempt,” is the source of the English word immunity, a state of protection from infectious disease. Perhaps the earliest written reference to the phenomenon of immunity can be traced back to Thucydides, the great historian of the Peloponnesian War. In describing a plague in Athens, he wrote in 430 bc that only those who had recovered from the plague could nurse the sick because they would not contract the disease a second time. Although early societies recognized the phenomenon of immunity, almost 2000 years passed before the concept was successfully converted into medically effective practice. Early Vaccination Studies Led the Way to Immunology The first recorded attempts to deliberately induce immunity were performed by the Chinese and Turks in the fifteenth century. They were attempting to prevent smallpox, a disease that is fatal in about 30% of cases and that leaves survivors disfigured for life (Figure 1-1). Reports suggest that the dried crusts derived from smallpox pustules were either inhaled or inserted into small cuts in the skin (a technique called variolation) in order to prevent this dreaded disease. In 1718, Lady Mary Wortley Montagu, the wife of the British ambassador in Constantinople, observed the positive effects of variolation on the native Turkish population and had the technique performed on her own children. The English physician Edward Jenner later made a giant advance in the deliberate development of immunity, again targeting smallpox. In 1798, intrigued by the fact that milkmaids who had contracted the mild disease cowpox were subsequently immune to the much more severe smallpox, Jenner reasoned that introducing fluid from a cowpox pustule into people (i.e., inoculating them) might protect them from smallpox. To test this idea, he inoculated an eight-year-old boy with fluid from a cowpox pustule and later intentionally infected the child with smallpox. As predicted, the child did not develop smallpox. Although this represented a major breakthrough, as one might imagine, these sorts of human studies could not be conducted under current standards of medical ethics. Jenner’s technique of inoculating with cowpox to protect against smallpox spread quickly through Europe. However, it FIGURE 1-1 African child with rash typical of smallpox on face, chest, and arms. Smallpox, caused by the virus Variola major, has a 30% mortality rate. Survivors are often left with disfiguring scars. [Centers for Disease Control.] was nearly a hundred years before this technique was applied to other diseases. As so often happens in science, serendipity combined with astute observation led to the next major advance in immunology: the induction of immunity to cholera. Louis Pasteur had succeeded in growing the bacterium that causes fowl cholera in culture, and confirmed this by injecting it into chickens that then developed fatal cholera. After returning from a summer vacation, he and colleagues resumed their experiments, injecting some chickens with an old bacterial culture. The chickens became ill, but to Pasteur’s surprise, they recovered. Interested, Pasteur then grew a fresh culture of the bacterium with the intention of injecting this lethal brew into some fresh, unexposed chickens. But as the story is told, his supply of fresh chickens was limited, and therefore he used a mixture of previously injected chickens and unexposed birds. Unexpectedly, only the fresh chickens died, while the chickens previously exposed to the older bacterial culture were completely protected from the disease. Pasteur hypothesized and later showed that aging had weakened the virulence of the pathogen and that such a weakened or attenuated strain could be administered to provide immunity against the disease. He called this attenuated strain a Overview of the Immune System vaccine (from the Latin vacca, meaning “cow”), in honor of Jenner’s work with cowpox inoculation. Pasteur extended these findings to other diseases, demonstrating that it was possible to attenuate a pathogen and administer the attenuated strain as a vaccine. In a now classic experiment performed in the small village of Pouilly-le-Fort in 1881, Pasteur first vaccinated one group of sheep with anthrax bacteria (Bacillus anthracis) that were attenuated by heat treatment. He then challenged the vaccinated sheep, along with some unvaccinated sheep, with a virulent culture of the anthrax bacillus. All the vaccinated sheep lived and all the unvaccinated animals died. These experiments marked the beginnings of the discipline of immunology. In 1885, Pasteur administered his first vaccine to a human, a young boy who had been bitten repeatedly by a rabid dog (Figure 1-2). The boy, Joseph Meister, was inoculated with a series of attenuated rabies virus preparations. The rabies vaccine is one of very few that can be successful when administered shortly after exposure, as long as the virus has not yet reached the central nervous system and begun to induce neurologic symptoms. Joseph lived, and later became a caretaker at the Pasteur Institute, which was opened in 1887 to treat the many rabies victims that began to flood in when word of Pasteur’s success spread; it remains to this day an institute dedicated to the prevention and treatment of infectious disease. FIGURE 1-2 Wood engraving of Louis Pasteur watching Joseph Meister receive the rabies vaccine. [Source: From Harper’s Weekly 29:836; courtesy of the National Library of Medicine.] | CHAPTER 1 3 Vaccination Is an Ongoing, Worldwide Enterprise The emergence of the study of immunology and the discovery of vaccines are tightly linked. The development of effective vaccines for some pathogens is still a major challenge, discussed in greater detail in Chapter 17. However, despite many biological and social hurdles, vaccination has yielded some of the most profound success stories in terms of improving mortality rates worldwide, especially in very young children. In 1977, the last known case of naturally acquired smallpox was seen in Somalia. This dreaded disease was eradicated by universal application of a vaccine similar to that used by Jenner in the 1790s. One consequence of eradication is that universal vaccination becomes unnecessary. This is a tremendous benefit, as most vaccines carry at least a slight risk to persons vaccinated. And yet in many cases every individual does not need to be immune in order to protect most of the population. As a critical mass of people acquire protective immunity, either through vaccination or infection, they can serve as a buffer for the rest. This principle, called herd immunity, works by decreasing the number of individuals who can harbor and spread an infectious agent, significantly decreasing the chances that susceptible individuals will become infected. This presents an important altruistic consideration: although many of us could survive infectious diseases for which we receive a vaccine (such as the flu), this is not true for everyone. Some individuals cannot receive the vaccine (e.g., the very young or immune compromised), and vaccination is never 100% effective. In other words, the susceptible, nonimmune individuals among us can benefit from the pervasive immunity of their neighbors. However, there is a darker side to eradication and the end of universal vaccination. Over time, the number of people with no immunity to the disease will begin to rise, ending herd immunity. Vaccination for smallpox largely ended by the early to mid-1970s, leaving well over half of the current world population susceptible to the disease. This means that smallpox, or a weaponized version, is now considered a potential bioterrorism threat. In response, new and safer vaccines against smallpox are still being developed today, most of which go toward vaccinating U.S. military personnel thought to be at greatest risk of possible exposure. In the United States and other industrialized nations, vaccines have eliminated a host of childhood diseases that were the cause of death for many young children just 50 years ago. Measles, mumps, chickenpox, whooping cough (pertussis), tetanus, diphtheria, and polio, once thought of as an inevitable part of childhood are now extremely rare or nonexistent in the United States because of current vaccination practices (Table 1-1). One can hardly estimate the savings to society resulting from the prevention of these diseases. Aside from suffering and mortality, the cost to treat these illnesses and their aftereffects or sequelae (such as paralysis, deafness, blindness, and mental retardation) is immense and dwarfs the costs of immunization. In fact, recent estimates suggest PA R T I 4 TABLE 1-1 | Introduction Cases of selected infectious disease in the United States before and after the introduction of effective vaccines Disease Smallpox ANNUAL CASES/YR CASES IN 2010 Prevaccine Postvaccine 48,164 Reduction (%) 0 100 100 Diphtheria 175,885 0 Rubeola (measles) 503,282 26 99.99 Mumps 152,209 2,612 98.28 Pertussis (“whooping cough”) 147,271 27,550 81.29 Paralytic polio 16,316 0 Rubella (German measles) 47,745 5 Tetanus (“lockjaw”) Invasive Haemophilus influenzae 1,314 (deaths) 20,000 26 (cases) 3,151 100 99.99 98.02 84.25 SOURCE: Adapted from W. A. Orenstein et al., 2005. Health Affairs 24:599 and CDC statistics of Notifiable Diseases. that significant economic and human life benefits could be realized by simply scaling up the use of a few childhood vaccines in the poorest nations, which currently bear the brunt of the impact of these childhood infectious diseases. For example, it is estimated that childhood pneumonia alone, caused primarily by vaccine-preventable Streptococcus pneumoniae (aka, pneumococcus) and Haemophilus influenzae type b (aka, Hib), will account for 2.7 million childhood deaths in developing nations over the next decade if vaccine strategies in these regions remain unchanged. Despite the many successes of vaccine programs, such as the eradication of smallpox, many vaccine challenges still remain. Perhaps the greatest current challenge is the design of effective vaccines for major killers such as malaria and AIDS. Using our increased understanding of the immune system, plus the tools of molecular and cellular biology, genomics, and proteomics, scientists will be better positioned to make progress toward preventing these and other emerging infectious diseases. A further issue is the fact that millions of children in developing countries die from diseases that are fully preventable by available, safe vaccines. High manufacturing costs, instability of the products, and cumbersome delivery problems keep these vaccines from reaching those who might benefit the most. This problem could be alleviated in many cases by development of future-generation vaccines that are inexpensive, heat stable, and administered without a needle. Finally, misinformation and myth surrounding vaccine efficacy and side effects continues to hamper many potentially life-saving vaccination programs (see Clinical Focus Box on p. 5). Immunology Is About More Than Just Vaccines and Infectious Disease For some diseases, immunization programs may be the best or even the only effective defense. At the top of this list are infec- tious diseases that can cause serious illness or even death in unvaccinated individuals, especially those transmitted by microbes that also spread rapidly between hosts. However, vaccination is not the only way to prevent or treat infectious disease. First and foremost is preventing infection, where access to clean water, good hygiene practices, and nutrient-rich diets can all inhibit transmission of infectious agents. Second, some infectious diseases are self-limiting, easily treatable, and nonlethal for most individuals, making them unlikely targets for costly vaccination programs. These include the common cold, caused by the Rhinovirus, and cold sores that result from Herpes Simplex Virus infection. Finally, some infectious agents are just not amenable to vaccination. This could be due to a range of factors, such as the number of different molecular variants of the organism, the complexity of the regimen required to generate protective immunity, or an inability to establish the needed immunologic memory responses (more on this later). One major breakthrough in the treatment of infectious disease came when the first antibiotics were introduced in the 1920s. Currently there are more than a hundred different antibiotics on the market, although most fall into just six or seven categories based on their mode of action. Antibiotics are chemical agents designed to destroy certain types of bacteria. They are ineffective against other types of infectious agents, as well as some bacterial species. One particularly worrying trend is the steady rise in antibiotic resistance among strains traditionally amenable to these drugs, making the design of next-generation antibiotics and new classes of drugs increasingly important. Although antiviral drugs are also available, most are not effective against many of the most common viruses, including influenza. This makes preventive vaccination the only real recourse against many debilitating infectious agents, even those that rarely cause mortality in healthy adults. For instance, because of the high mutation rate of the influenza virus, each year a new flu vaccine must Overview of the Immune System | CHAPTER 1 5 BOX 1-1 CLINICAL FOCUS Vaccine Controversy: What’s Truth and What’s Myth? Despite the record of worldwide success of vaccines in improving public health, some opponents claim that vaccines do more harm than good, pressing for elimination or curtailment of childhood vaccination programs. There is no dispute that vaccines represent unique safety issues, since they are administered to people who are healthy. Furthermore, there is general agreement that vaccines must be rigorously tested and regulated, and that the public must have access to clear and complete information about them. Although the claims of vaccine critics must be evaluated, many can be answered by careful and objective examination of records. A recent example is the claim that vaccines given to infants and very young children may contribute to the rising incidence of autism. This began with the suggestion that thimerosal, a mercury-based additive used to inhibit bacterial growth in some vaccine preparations since the 1930s, was causing autism in children. In 1999 the U.S. Centers for Disease Control and Prevention (CDC) and the American Association of Pediatricians (AAP) released a joint recommendation that vaccine manufacturers begin to gradually phase out thimerosal use in vaccines. This recommendation was based on the increase in the number of vaccines given to infants and was aimed at keeping children at or below Environmental Protection Agency (EPA)–recommended maximums in mercury exposure. However, with the release of this recommendation, parent-led public advocacy groups began a media-fueled campaign to build a case demonstrating what they believed was a link between vaccines and an epidemic of autism. These AAP recommendations and public fears led to a dramatic decline in the latter half of 1999 in U.S. newborns vaccinated for hepatitis B. To date, no credible study has shown a scientific link between thimerosal and autism. In fact, cases of autism in children have continued to rise since thimerosal was removed from all childhood vaccines in 2001. Despite evidence to the contrary, some still believe this claim. A 1998 study appearing in The Lancet, a reputable British medical journal, further fueled these parent advocacy groups and anti-vaccine organizations. The article, published by Andrew Wakefield, claimed the measles-mumps-rubella (MMR) vaccine caused pervasive developmental disorders in children, including autism spectrum disorder. More than a decade of subsequent research has been unable to substantiate these claims, and 10 of the original 12 authors on the paper later withdrew their support for the conclusions of the study. In 2010, The Lancet retracted the original article when it was shown that the data in the study had been falsified to reach desired conclusions. Nonetheless, in the years between the original publication of the Lancet article and its retraction, this case is credited with decreasing rates of MMR vaccination from a high of 92% to a low of almost 60% in certain areas of the United Kingdom. The resulting expansion in the population of susceptible individuals led to endemic rates of measles and mumps infection, especially in several areas of Europe, and is credited with thou- be prepared based on a prediction of the prominent genotypes likely to be encountered in the next season. Some years this vaccine is more effective than others. If and when a more lethal and unexpected pandemic strain arises, there will be a race between its spread and the manufacture and administration of a new vaccine. With the current ease of worldwide travel, present-day emergence of a pandemic strain of influenza could dwarf the devastation wrought by the 1918 flu pandemic, which left up to 50 million dead. sands of extended hospitalizations and several deaths in infected children. Why has there been such a strong urge to cling to the belief that childhood vaccines are linked with developmental disorders in children despite much scientific evidence to the contrary? One possibility lies in the timing of the two events. Based on current AAP recommendations, in the United States most children receive 14 different vaccines and a total of up to 26 shots by the age of 2. In 1983, children received less than half this number of vaccinations. Couple this with the onset of the first signs of autism and other developmental disorders in children, which can appear quite suddenly and peak around 2 years of age. This sharp rise in the number of vaccinations young children receive today and coincidence in timing of initial autism symptoms is credited with sparking these fears about childhood vaccines. Add to this the increasing drop in basic scientific literacy by the general public and the overabundance of ways to gather such information (accurate or not). As concerned parents search for answers, one can begin to see how even scientifically unsupported links could begin to take hold as families grapple with how to make intelligent public health risk assessments. The notion that vaccines cause autism was rejected long ago by most scientists. Despite this, more work clearly needs to be done to bridge the gap between public perception and scientific understanding. Gross, L. 2009. A broken trust: Lessons from the vaccine–autism wars. PLoS Biology 7:e1000114. Larson, H.J., et al. 2011. Addressing the vaccine confidence gap. Lancet 378:526. However, the eradication of infectious disease is not the only worthy goal of immunology research. As we will see later, exposure to infectious agents is part of our evolutionary history, and wiping out all of these creatures could potentially cause more harm than good, both for the host and the environment. Thanks to many technical advances allowing scientific discoveries to move efficiently from the bench to the bedside, clinicians can now manipulate the immune response in ways never before possible. For example, treatments to boost, inhibit, 6 PA R T I | Introduction or redirect the specific efforts of immune cells are being applied to treat autoimmune disease, cancer, and allergy, as well as other chronic disorders. These efforts are already extending and saving lives. Likewise, a clearer understanding of immunity has highlighted the interconnected nature of body systems, providing unique insights into areas such as cell biology, human genetics, and metabolism. While a cure for AIDS and a vaccine to prevent HIV infection are still the primary targets for many scientists who study this disease, a great deal of basic science knowledge has been gleaned from the study of just this one virus and its interaction with the human immune system. TABLE 1-2 Immunity Involves Both Humoral and Cellular Components Pasteur showed that vaccination worked, but he did not understand how. Some scientists believed that immune protection in vaccinated individuals was mediated by cells, while others postulated that a soluble agent delivered protection. The experimental work of Emil von Behring and Shibasaburo Kitasato in 1890 gave the first insights into the mechanism of immunity, earning von Behring the Nobel Prize in Physiology or Medicine in 1901 (Table 1-2). Von Behring and Nobel Prizes for immunologic research Year Recipient Country Research 1901 Emil von Behring Germany Serum antitoxins 1905 Robert Koch Germany Cellular immunity to tuberculosis 1908 Elie Metchnikoff Paul Ehrlich Russia Germany Role of phagocytosis (Metchnikoff ) and antitoxins (Ehrlich) in immunity 1913 Charles Richet France Anaphylaxis 1919 Jules Bordet Belgium Complement-mediated bacteriolysis 1930 Karl Landsteiner United States Discovery of human blood groups 1951 Max Theiler South Africa Development of yellow fever vaccine 1957 Daniel Bovet Switzerland Antihistamines 1960 F. Macfarlane Burnet Peter Medawar Australia Great Britain Discovery of acquired immunological tolerance 1972 Rodney R. Porter Gerald M. Edelman Great Britain United States Chemical structure of antibodies 1977 Rosalyn R. Yalow United States Development of radioimmunoassay 1980 George Snell Jean Dausset Baruj Benacerraf United States France United States Major histocompatibility complex 1984 Niels K. Jerne Cesar Milstein Georges E. Köhler Denmark Great Britain Germany Immune regulatory theories (Jerne) and technological advances in the development of monoclonal antibodies (Milstein and Köhler) 1987 Susumu Tonegawa Japan Gene rearrangement in antibody production 1991 E. Donnall Thomas Joseph Murray United States United States Transplantation immunology 1996 Peter C. Doherty Rolf M. Zinkernagel Australia Switzerland Role of major histocompatibility complex in antigen recognition by T cells 2002 Sydney Brenner H. Robert Horvitz J. E. Sulston South Africa United States Great Britain Genetic regulation of organ development and cell death (apoptosis) 2008 Harald zur Hausen Françoise Barré-Sinoussi Luc Montagnier Germany France France Role of HPV in causing cervical cancer (Hausen) and the discovery of HIV (Barré-Sinoussi and Montagnier) 2011 Jules Hoffman Bruce Beutler Ralph Steinman France United States United States Discovery of activating principles of innate immunity (Hoffman and Beutler) and role of dendritic cells in adaptive immunity (Steinman) Overview of the Immune System | CHAPTER 1 7 FIGURE 1-3 Drawing by Elie Metchnikoff of phagocytic cells surrounding a foreign particle (left) and modern image of a phagocyte engulfing the bacteria that cause tuberculosis (right). Metchnikoff first described and named the process of phagocytosis, or ingestion of foreign matter by white blood cells. Today, phagocytic cells can be imaged in great detail using advanced microscopy techniques. [Drawing reproduced by permission of The British Library:7616.h.19, Lectures on the Comparative Pathology of Inflammation delivered at the Pasteur Institute in 1891, translated by F. A. Starling and E. H. Starling, with plates by II’ya II’ich Mechnikov, 1893, p. 64, fig. 32. Photo courtesy Dr. Volker Brinkmann/Visuals Unlimited, Inc.] Kitasato demonstrated that serum—the liquid, noncellular component recovered from coagulated blood—from animals previously immunized with diphtheria could transfer the immune state to unimmunized animals. In 1883, even before the discovery that a serum component could transfer immunity, Elie Metchnikoff, another Nobel Prize winner, demonstrated that cells also contribute to the immune state of an animal. He observed that certain white blood cells, which he termed phagocytes, ingested (phagocytosed) microorganisms and other foreign material (Figure 1-3, left). Noting that these phagocytic cells were more active in animals that had been immunized, Metchnikoff hypothesized that cells, rather than serum components, were the major effectors of immunity. The active phagocytic cells identified by Metchnikoff were likely blood monocytes and neutrophils (see Chapter 2), which can now be imaged using very sophisticated microscopic techniques (Figure 1-3, right). Humoral Immunity The debate over cells versus soluble mediators of immunity raged for decades. In search of the protective agent of immunity, various researchers in the early 1900s helped characterize the active immune component in blood serum. This soluble component could neutralize or precipitate toxins and could agglutinate (clump) bacteria. In each case, the component was named for the activity it exhibited: antitoxin, precipitin, and agglutinin, respectively. Initially, different serum components were thought to be responsible for each activity, but during the 1930s, mainly through the efforts of Elvin Kabat, a fraction of serum first called gamma globulin (now immunoglobulin) was shown to be responsible for all these activities. The soluble active molecules in the immunoglobulin fraction of serum are now commonly referred to as antibodies. Because these antibodies were contained in body fluids (known at that time as the body humors), the immunologic events they participated in was called humoral immunity. The observation of von Behring and Kitasato was quickly applied to clinical practice. Antiserum, the antibodycontaining serum fraction from a pathogen-exposed individual, derived in this case from horses, was given to patients suffering from diphtheria and tetanus. A dramatic vignette of this application is described in the Clinical Focus box on page 8. Today there are still therapies that rely on transfer of immunoglobulins to protect susceptible individuals. For example, emergency use of immune serum, containing antibodies against snake or scorpion venoms, is a common practice for treating bite victims. This form of immune protection that is transferred between individuals is called passive immunity because the individual receiving it did not make his or her own immune response against the pathogen. Newborn infants benefit from passive immunity by the presence of maternal antibodies in their circulation. Passive immunity may also be used as a preventive (prophylaxis) to boost the immune potential of those with compromised immunity or who anticipate future exposure to a particular microbe. While passive immunity can supply a quick solution, it is short-lived and limited, as the cells that produce these antibodies are not being transferred. On the other hand, administration 8 PA R T I | Introduction CLINICAL FOCUS Passive Antibodies and the Iditarod In 1890, immunologists Emil Behring and Shibasaburo Kitasato, working together in Berlin, reported an extraordinary experiment. After immunizing rabbits with tetanus and then collecting blood serum from these animals, they injected a small amount of immune serum (cell-free fluid) into the abdominal cavity of six mice. Twenty-four hours later, they infected the treated mice and untreated controls with live, virulent tetanus bacteria. All of the control mice died within 48 hours of infection, whereas the treated mice not only survived but showed no effects of infection. This landmark experiment demonstrated two important points. One, it showed that substances that could protect an animal against pathogens appeared in serum following immunization. Two, this work demonstrated that immunity could be passively acquired, or transferred from one animal to another by taking serum from an immune animal and injecting it into a nonimmune one. These and subsequent experiments did not go unnoticed. Both men eventually received titles (Behring became von Behring and Kitasato became Baron Kitasato). A few years later, in 1901, von Behring was awarded the first Nobel Prize in Physiology or Medicine (see Table 1-2). These early observations, and others, paved the way for the introduction of passive immunization into clinical practice. During the 1930s and 1940s, passive immunotherapy, the endowment of resistance to pathogens by transfer of antibodies from an immunized donor to an unimmunized recipient, was used to prevent or modify the course of measles and hepatitis A. Subsequently, clinical experience and advances in the technology of immunoglobulin preparation have made this approach a standard medical practice. Passive immunization based on the transfer of antibodies is widely used in the treatment of immunodeficiency and some autoimmune diseases. It is also used to protect individuals against anticipated exposure to infectious and toxic agents against which they have no immunity. Finally, passive immunization can be lifesaving during episodes of certain types of acute infection, such as following exposure to rabies virus. Immunoglobulin for passive immunization is prepared from the pooled of a vaccine or natural infection is said to engender active immunity in the host: the production of one’s own immunity. The induction of active immunity can supply the individual with a renewable, long-lived protection from the specific infectious organism. As we discuss further below, this long-lived protection comes from memory cells, which provide protection for years or even decades after the initial exposure. Cell-Mediated Immunity A controversy developed between those who held to the concept of humoral immunity and those who agreed with Metchnikoff ’s concept of immunity imparted by specific cells, or cell-mediated immunity. The relative contributions of the two were widely debated at the time. It is now obvious that both are correct—the full immune response requires both cellular and humoral (soluble) components. Early studies of immune cells were hindered by the lack of genetically defined animal models and modern tissue culture tech- plasma of thousands of donors. In effect, recipients of these antibody preparations are receiving a sample of the antibodies produced by many people to a broad diversity of pathogens—a gram of intravenous immune globulin (IVIG) contains about 1018 molecules of antibody and recognize more than 107 different antigens. A product derived from the blood of such a large number of donors carries a risk of harboring pathogenic agents, particularly viruses. This risk is minimized by modern-day production techniques. The manufacture of IVIG involves treatment with solvents, such as ethanol, and the use of detergents that are highly effective in inactivating viruses such as HIV and hepatitis. In addition to treatment against infectious disease, or acute situations, IVIG is also used today for treating some chronic diseases, including several forms of immune deficiency. In all cases, the transfer of passive immunity supplies only temporary protection. One of the most famous instances of passive antibody therapy occurred in 1925, when an outbreak of diphtheria niques, whereas early studies with serum took advantage of the ready availability of blood and established biochemical techniques to purify proteins. Information about cellular immunity therefore lagged behind a characterization of humoral immunity. In a key experiment in the 1940s, Merrill Chase, working at The Rockefeller Institute, succeeded in conferring immunity against tuberculosis by transferring white blood cells between guinea pigs. Until that point, attempts to develop an effective vaccine or antibody therapy against tuberculosis had met with failure. Thus, Chase’s demonstration helped to rekindle interest in cellular immunity. With the emergence of improved cell culture and transfer techniques in the 1950s, the lymphocyte was identified as the cell type responsible for both cellular and humoral immunity. Soon thereafter, experiments with chickens pioneered by Bruce Glick at Mississippi State University indicated the existence of two types of lymphocytes: T lymphocytes (T cells), derived from Overview of the Immune System | CHAPTER 1 9 BOX 1-2 White Mountain Golovin Koyuk Galena Nulato Nome Ruby Safety Elim Kaktag Cripple Shaktoolik Nenana Unalakleet Nikolai Ophir Takotna Rohn McGrath Skwentna Rainy Pass Yentna Finger Lake Willow Anchorage Campell Airstrip FIGURE 1 (left) Leonhard Seppala, the Norwegian who led a team of sled dogs in the 1925 diphtheria antibody run from Nenana to Nome, Alaska. (right) Map of the current route of the Iditarod Race, which commemorates this historic delivery of lifesaving antibody. [Source: Underwood & Underwood/Corbis.] was diagnosed in what was then the remote outpost of Nome, Alaska. Lifesaving diphtheria-specific antibodies were available in Anchorage, but no roads were open and the weather was too dangerous for flight. History tells us that 20 mushers set up a dogsled relay to cover the almost 700 miles between Nenana, the end of the railroad run, and remote Nome. In this relay, two Norwegians and their dogs covered particularly critical territory and withstood blizzard conditions: Leonhard Seppala (Figure 1, left), who covered the most treacherous territory, and Gunnar Kaasen, who drove the final two legs in whiteout conditions, behind his lead dog Balto. Kaasen and Balto arrived in time to save many of the chil- the thymus, and B lymphocytes (B cells), derived from the bursa of Fabricius in birds (an outgrowth of the cloaca). In a convenient twist of nomenclature that makes B and T cell origins easier to remember, the mammalian equivalent of the bursa of Fabricius is bone marrow, the home of developing B cells in mammals. We now know that cellular immunity is imparted by T cells and that the antibodies produced by B cells confer humoral immunity. The real controversy about the roles of humoral versus cellular immunity was resolved when the two systems were shown to be intertwined and it became clear that both are necessary for a complete immune response against most pathogens. How Are Foreign Substances Recognized by the Immune System? One of the great enigmas confronting early immunologists was what determines the specificity of the immune response dren in the town. To commemorate this heroic event, later that same year a statue of Balto was placed in Central Park, New York City, where it still stands today. This journey is memorialized every year in the running of the Iditarod sled dog race. A map showing the current route of this more than 1000-mile trek is shown in Figure 1, right. for a particular foreign material, or antigen, the general term for any substance that elicits a specific response by B or T lymphocytes. Around 1900, Jules Bordet at the Pasteur Institute expanded the concept of immunity beyond infectious diseases, demonstrating that nonpathogenic substances, such as red blood cells from other species, could also serve as antigens. Serum from an animal that had been inoculated with noninfectious but otherwise foreign (nonself) material would nevertheless react with the injected material in a specific manner. The work of Karl Landsteiner and those who followed him showed that injecting an animal with almost any nonself organic chemical could induce production of antibodies that would bind specifically to the chemical. These studies demonstrated that antibodies have a capacity for an almost unlimited range of reactivity, including responses to compounds that had only recently been synthesized in the laboratory and are otherwise not found in nature! In addition, it was 10 PA R T I | Introduction shown that molecules differing in the smallest detail, such as a single amino acid, could be distinguished by their reactivity with different antibodies. Two major theories were proposed to account for this specificity: the selective theory and the instructional theory. The earliest conception of the selective theory dates to Paul Ehrlich in 1900. In an attempt to explain the origin of serum antibody, Ehrlich proposed that cells in the blood expressed a variety of receptors, which he called side-chain receptors, that could bind to infectious agents and inactivate them. Borrowing a concept used by Emil Fischer in 1894 to explain the interaction between an enzyme and its substrate, Ehrlich proposed that binding of the receptor to an infectious agent was like the fit between a lock and key. Ehrlich suggested that interaction between an infectious agent and a cell-bound receptor would induce the cell to produce and release more receptors with the same specificity (Figure 1-4). In Ehrlich’s mind, the cells were pluripotent, expressing a number of FIGURE 1-4 Representation of Paul Ehrlich’s side chain theory to explain antibody formation. In Ehrlich’s initial theory, the cell is pluripotent in that it expresses a number of different receptors or side chains, all with different specificities. If an antigen encounters this cell and has a good fit with one of its side chains, synthesis of that receptor is triggered and the receptor will be released. [From Ehrlich’s Croonian lecture of 1900 to the Royal Society.] different receptors, each of which could be individually “selected.” According to Ehrlich’s theory, the specificity of the receptor was determined in the host before its exposure to the foreign antigen, and therefore the antigen selected the appropriate receptor. Ultimately, most aspects of Ehrlich’s theory would be proven correct, with the following minor refinement: instead of one cell making many receptors, each cell makes many copies of just one membrane-bound receptor (one specificity). An army of cells, each with a different antigen specificity, is therefore required. The selected B cell can be triggered to proliferate and to secrete many copies of these receptors in soluble form (now called antibodies) once it has been selected by antigen binding. In the 1930s and 1940s, the selective theory was challenged by various instructional theories. These theories held that antigen played a central role in determining the specificity of the antibody molecule. According to the instructional theorists, a particular antigen would serve as a template around which antibody would fold—sort of like an impression mold. The antibody molecule would thereby assume a configuration complementary to that of the antigen template. This concept was first postulated by Friedrich Breinl and Felix Haurowitz in about 1930 and redefined in the 1940s in terms of protein folding by Linus Pauling. In the 1950s, selective theories resurfaced as a result of new experimental data. Through the insights of F. Macfarlane Burnet, Niels Jerne, and David Talmadge, this model was refined into a hypothesis that came to be known as the clonal selection theory. This hypothesis has been further refined and is now accepted as an underlying paradigm of modern immunology. According to this theory, an individual B or T lymphocyte expresses many copies of a membrane receptor that is specific for a single, distinct antigen. This unique receptor specificity is determined in the lymphocyte before it is exposed to the antigen. Binding of antigen to its specific receptor activates the cell, causing it to proliferate into a clone of daughter cells that have the same receptor specificity as the parent cell. The instructional theories were formally disproved in the 1960s, by which time information was emerging about the structure of protein, RNA, and DNA that would offer new insights into the vexing problem of how an individual could make antibodies against almost anything, sight unseen. Overview Figure 1-5 presents a very basic scheme of clonal selection in the humoral (B cell) and cellular (T cell) branches of immunity. We now know that B cells produce antibodies, a soluble version of their receptor protein, which bind to foreign proteins, flagging them for destruction. T cells, which come in several different forms, also use their surfacebound T-cell receptors to sense antigen. These cells can perform a range of different functions once selected by antigen encounter, including the secretion of soluble compounds to aid other white blood cells (such as B lymphocytes) and the destruction of infected host cells. Overview of the Immune System | CHAPTER 1 11 1-5 OVERVIEW FIGURE An Outline for the Humoral and Cell-Mediated (Cellular) Branches of the Immune System Foreign proteins or infectious agents Vertebrate body Humoral response (B lymphocytes) Cell-mediated response (T lymphocytes) T-cell receptor B cell B T-cell receptor + T B-cell receptor T + Antigen + Antigen Antigen T Antigen-selected antibodysecreting B cell T Killing of infected cells B Antibody Antigenselected T cells Cytokine secretion Antigen elimination The humoral response involves interaction of B cells with foreign proteins, called antigens, and their differentiation into antibodysecreting cells. The secreted antibody binds to foreign proteins or infectious agents, helping to clear them from the body. The cell- Important Concepts for Understanding the Mammalian Immune Response Today, more than ever, we are beginning to understand on a molecular and cellular level how a vaccine or infection leads to the development of immunity. As highlighted by mediated response involves various subpopulations of T lymphocytes, which can perform many functions, including the secretion of soluble messengers that help direct other cells of the immune system and direct killing of infected cells. the historical studies described above, this involves a complex system of cells and soluble compounds that have evolved to protect us against an enormous range of invaders of all shapes, sizes, and chemical structures. In this section, we cover the range of organisms that challenge the immune system and several of the important new concepts that are unique hallmarks of how the immune system carries out this task. 12 PA R T I | Introduction Pathogens Come in Many Forms and Must First Breach Natural Barriers Organisms causing disease are termed pathogens, and the process by which they induce illness in the host is called pathogenesis. The human pathogens can be grouped into four major categories based on shared characteristics: viruses, fungi, parasites, and bacteria (Table 1-3). Some example organisms from each category can be found in Figure 1-6. As we will see in the next section, some of the shared characteristics that are common to groups of pathogens, but not to the host, can be exploited by the immune system for recognition and destruction. The microenvironment in which the immune response begins to emerge can also influence the outcome; the same pathogen may be treated differently depending on the context in which it is encountered. Some areas of the body, such as the central nervous system, are virtually “off limits” for the immune system because the immune response could do more damage than the pathogen. In other cases, the environment may come with inherent directional cues for immune cells. For instance, some foreign compounds that enter via the digestive tract, including the commensal microbes that help us digest food, are tolerated by the immune system. However, when these same foreigners enter the bloodstream they are typically treated much more aggressively. Each encounter with pathogen thus engages a distinct set of strategies that depends on the nature of the invader and on the microenvironment in which engagement occurs. It is worth noting that immune pathways do not become engaged until foreign organisms first breach the physical barriers of the body. Obvious barriers include the skin and TABLE 1-3 the mucous membranes. The acidity of the stomach contents, of the vagina, and of perspiration poses a further barrier to many organisms, which are unable to grow in low pH conditions. The importance of these barriers becomes obvious when they are surmounted. Animal bites can communicate rabies or tetanus, whereas insect puncture wounds can transmit the causative agents of such diseases as malaria (mosquitoes), plague (fleas), and Lyme disease (ticks). A dramatic example is seen in burn victims, who lose the protective skin at the burn site and must be treated aggressively with drugs to prevent the rampant bacterial and fungal infections that often follow. The Immune Response Quickly Becomes Tailored to Suit the Assault With the above in mind, an effective defense relies heavily on the nature of the invading pathogen offense. The cells and molecules that become activated in a given immune response depend on the chemical structures present on the pathogen, whether it resides inside or outside of host cells, and the location of the response. This means that different chemical structures and microenvironmental cues need to be detected and appropriately evaluated, initiating the most effective response strategy. The process of pathogen recognition involves an interaction between the foreign organism and a recognition molecule (or molecules) expressed by host cells. Although these recognition molecules are frequently membranebound receptors, soluble receptors or secreted recognition molecules can also be engaged. Ligands for these recognition molecules can include whole pathogens, antigenic fragments of pathogens, or products secreted by these foreign organisms. The outcome of this ligand binding is an intracellular Major categories of human pathogens Major groups of human pathogens Specific examples Disease Viruses Poliovirus Variola Virus Human Immunodeficiency Virus Rubeola Virus Poliomyelitis (Polio) Smallpox AIDS Measles Fungi Candida albicans Tinea corporis Cryptococcus neoformans Candidiasis (Thrush) Ringworm Cryptococcal meningitis Parasites Plasmodium species Leishmania major Entamoeba histolytica Malaria Leishmaniasis Amoebic colitis Bacteria Mycobacterium tuberculosis Bordetella pertussis Vibrio cholerae Borrelia burgdorferi Tuberculosis Whooping cough (pertussis) Cholera Lyme disease Overview of the Immune System | CHAPTER 1 13 (a) Virus: Rotavirus (b) Fungus: Candida albicans (c) Parasite: Filaria (d) Bacterium: Mycobacterium tuberculosis FIGURE 1-6 Pathogens representing the major categories of microorganisms causing human disease. (a) Viruses: Transmission electron micrograph of rotavirus, a major cause of infant diarrhea. Rotavirus accounts for approximately 1 million infant deaths per year in developing countries and hospitalization of about 50,000 infants per year in the United States. (b) Fungi: Candida albicans, a yeast inhabiting human mouth, throat, intestines, and genitourinary tract; C. albicans commonly causes an oral rash (thrush) or vaginitis in immunosuppressed individuals or in those taking antibiotics that kill normal bacterial flora. (c) Parasites: The larval form of filaria, a parasitic worm, being attacked by macrophages. Approximately 120 million persons worldwide have some form of filariasis. (d) Bacteria: Mycobacterium tuberculosis, the bacterium that causes tuberculosis, being ingested by a human macrophage. [(a) Dr. Gary Gaugler/Getty Images; or extracellular cascade of events that ultimately leads to the labeling and destruction of the pathogen—simply referred to as the immune response. The entirety of this response is actually engagement of a complex system of cells that can recognize and kill or engulf a pathogen (cellular immunity), as well as a myriad of soluble proteins that help to orchestrate labeling and destruction of foreign invaders (humoral immunity). The nature of the immune response will vary depending on the number and type of recognition molecules engaged. For instance, all viruses are tiny, obligate, intracellular patho- (b) SPL/Photo Researchers; (c) Oliver Meckes/Nicole Ottawa/Eye of Science/ Photo Researchers; (d) Max Planck Institute for Infection Biology/Dr. Volker Brinkmann.] gens that spend the majority of their life cycle residing inside host cells. An effective defense strategy must therefore involve identification of infected host cells along with recognition of the surface of the pathogen. This means that some immune cells must be capable of detecting changes that occur in a host cell after it becomes infected. This is achieved by a range of cytotoxic cells but especially cytotoxic T lymphocytes (aka CTLs, or Tc cells), a part of the cellular arm of immunity. In this case, recognition molecules positioned inside cells are key to the initial response. These intracellular receptors bind to viral proteins present in the cytosol and 14 PA R T I | Introduction initiate an early warning system, alerting the cell to the presence of an invader. Sacrifice of virally infected cells often becomes the only way to truly eradicate this type of pathogen. In general, this sacrifice is for the good of the whole organism, although in some instances it can cause disruptions to normal function. For example, the Human Immunodeficiency Virus (HIV) infects a type of T cell called a T helper cell (TH cell). These cells are called helpers because they guide the behavior of other immune cells, including B cells, and are therefore pivotal for selecting the pathway taken by the immune response. Once too many of these cells are destroyed or otherwise rendered nonfunctional, many of the directional cues needed for a healthy immune response are missing and fighting all types of infections becomes problematic. As we discuss later in this chapter, the resulting immunodeficiency allows opportunistic infections to take hold and potentially kill the patient. Similar but distinct immune mechanisms are deployed to mediate the discovery of extracellular pathogens, such as fungi, most bacteria, and some parasites. These rely primarily on cell surface or soluble recognition molecules that probe the extracellular spaces of the body. In this case, B cells and the antibodies they produce as a part of humoral immunity play major roles. For instance, antibodies can squeeze into spaces in the body where B cells themselves may not be able to reach, helping to identify pathogens hiding in these out-of-reach places. Large parasites present yet another problem; they are too big for phagocytic cells to envelop. In this case, cells that can deposit toxic substances or that can secrete products that induce expulsion (e.g., sneezing, coughing, vomiting) become a better strategy. As we study the complexities of the mammalian immune response, it is worth remembering that a single solution does not exist for all pathogens. At the same time, these various immune pathways carry out their jobs with considerable overlap in structure and in function. Pathogen Recognition Molecules Can Be Encoded in the Germline or Randomly Generated As one might imagine, most pathogens express at least a few chemical structures that are not typically found in mammals. Pathogen-associated molecular patterns (or PAMPs) are common foreign structures that characterize whole groups of pathogens. It is these unique antigenic structures that the immune system frequently recognizes first. Animals, both invertebrates and vertebrates, have evolved to express several types of cell surface and soluble proteins that quickly recognize many of these PAMPs; a form of pathogen profiling. For example, encapsulated bacteria possess a polysaccharide coat with a unique chemical structure that is not found on other bacterial or human cells. White blood cells naturally express a variety of receptors, collectively referred to as pattern recognition receptors (PRRs), that specifically recognize these sugar residues, as well as other common foreign struc- tures. When PRRs detect these chemical structures, a cascade of events labels the target pathogen for destruction. PPRs are proteins encoded in the genomic DNA and are always expressed by many different immune cells. These conserved, germline-encoded recognition molecules are thus a first line of defense for the quick detection of many of the typical chemical identifiers carried by the most common invaders. A significant and powerful corollary to this is that it allows early categorizing or profiling of the sort of pathogen of concern. This is key to the subsequent immune response routes that will be followed, and therefore the fine tailoring of the immune response as it develops. For example, viruses frequently expose unique chemical structures only during their replication inside host cells. Many of these can be detected via intracellular receptors that bind exposed chemical moieties while still inside the host cell. This can trigger an immediate antiviral response in the infected cell that blocks further virus replication. At the same time, this initiates the secretion of chemical warning signals sent to nearby cells to help them guard against infection (a neighborhood watch system!). This early categorizing happens via a subtle tracking system that allows the immune response to make note of which recognition molecules were involved in the initial detection event and relay that information to subsequent responding immune cells, allowing the follow-up response to begin to focus attention on the likely type of assault underway. Host-pathogen interactions are an ongoing arms race; pathogens evolve to express unique structures that avoid host detection, and the host germline-encoded recognition system co-evolves to match these new challenges. However, because pathogens generally have much shorter life cycles than their vertebrate hosts, and some utilize error-prone DNA polymerases to replicate their genomes, pathogens can evolve rapidly to evade host encoded recognition systems. If this were our only defense, the host immune response would quickly become obsolete thanks to these real-time pathogen avoidance strategies. How can the immune system prepare for this? How can our DNA encode a recognition system for things that change in random ways over time? Better yet, how do we build a system to recognize new chemical structures that may arise in the future? Thankfully, the vertebrate immune system has evolved a clever, albeit resource intensive, response to this dilemma: to favor randomness in the design of some recognition molecules. This strategy, called generation of diversity, is employed only by developing B and T lymphocytes. The result is a group of B and T cells where each expresses many copies of one unique recognition molecule, resulting in a population with the theoretical potential to respond to any antigen that may come along (Figure 1-7). This feat is accomplished by rearranging and editing the genomic DNA that encodes the antigen receptors expressed by each B or T lymphocyte. Not unlike the error-prone DNA replication method employed by pathogens, this system allows chance to play a role in generating a menu of responding recognition molecules. As one might imagine, however, this cutting and splicing of chromosomes is not without risk. Many B and T cells do not | Overview of the Immune System Generation of diversity Deletion CHAPTER 1 15 Clonal selection and expansion 2 1 2 Antigen 2 2 2 2 Stem cell 3 2 3 2 4 Primary lymphoid organs FIGURE 1-7 Generation of diversity and clonal selection in T and B lymphocytes. Maturation in T and B cells, which occurs in primary lymphoid organs (bone marrow for B cells and thymus for T cells) in the absence of antigen, produces cells with a committed antigenic specificity, each of which expresses many copies of surface receptor that binds to one particular antigen. Different clones of B cells (1, 2, 3, and 4) are illustrated in this figure. Cells that do not die or become deleted during this maturation and weeding-out process move into the circulation of the body and are available to interact with antigen. There, clonal selection occurs when one of these cells survive this DNA surgery or the quality control processes that follow, all of which take place in primary lymphoid organs: the thymus for T cells and bone marrow for B cells. Surviving cells move into the circulation of the body, where they are available if their specific, or cognate, antigen is encountered. When antigens bind to the surface receptors on these cells, they trigger clonal selection (see Figure 1-7). The ensuing proliferation of the selected clone of cells creates an army of cells, all with the same receptor and responsible for binding more of the same antigen, with the ultimate goal of destroying the pathogen in question. In B lymphocytes, these recognition molecules are B-cell receptors when they are surface structures and antibodies in their secreted form. In T lymphocytes, where no soluble form exists, they are T-cell receptors. In 1976 Susumu Tonegawa, then at The Basel Institute for Immunology in Switzerland, discovered the molecular mechanism behind the DNA recombination events that generate B-cell receptors and antibodies (Chapter 7 covers this in detail). This Circulation through the body encounters its cognate or specific antigen. Clonal proliferation of an antigen-activated cell (number 2 or pink in this example) leads to many cells that can engage with and destroy the antigen, plus memory cells that can be called upon during a subsequent exposure. The B cells secrete antibody, a soluble form of the receptor, reactive with the activating antigen. Similar processes take place in the T-lymphocyte population, resulting in clones of memory T cells and effector T cells; the latter include activated TH cells, which secrete cytokines that aid in the further development of adaptive immunity, and cytotoxic T lymphocytes (CTLs), which can kill infected host cells. was a true turning point in immunologic understanding; for this discovery he received widespread recognition, including the 1987 Nobel Prize in Physiology or Medicine (see Table 1-2). Tolerance Ensures That the Immune System Avoids Destroying the Host One consequence of generating random recognition receptors is that some could recognize and target the host. In order for this strategy to work effectively, the immune system must somehow avoid accidentally recognizing and destroying host tissues. This principle, which relies on self/ nonself discrimination, is called tolerance, another hallmark of the immune response. The work credited with its illumination also resulted in a Nobel Prize in Physiology or Medicine, awarded to F. Macfarlane Burnet and Peter Medawar in 1960. Burnet was the first to propose that exposure to nonself antigens during certain stages of life could result in an 16 PA R T I | Introduction immune system that ignored these antigens later. Medawar later proved the validity of this theory by exposing mouse embryos to foreign antigens and showing that these mice developed the ability to tolerate these antigens later in life. To establish tolerance, the antigen receptors present on developing B and T cells must first pass a test of nonresponsiveness against host structures. This process, which begins shortly after these randomly generated receptors are produced, is achieved by the destruction or inhibition of any cells that have inadvertently generated receptors with the ability to harm the host. Successful maintenance of tolerance ensures that the host always knows the difference between self and nonself (usually referred to as foreign). One recent re-envisioning of how tolerance is operationally maintained is called the danger hypothesis. This hypothesis suggests that the immune system constantly evaluates each new encounter more for its potential to be dangerous to the host than for whether it is self or not. For instance, cell death can have many causes, including natural homeostatic processes, mechanical damage, or infection. The former is a normal part of the everyday biological events in the body, and only requires a cleanup response to remove debris. The latter two, however, come with warning signs that include the release of intracellular contents, expression of cellular stress proteins, or pathogen-specific products. These pathogen or cell-associated stress compounds, sometimes referred to as danger signals, can engage specific host recognition molecules (e.g., PRRs) that deliver a signal to immune cells to get involved during these unnatural causes of cellular death. One unintended consequence of robust self-tolerance is that the immune system frequently ignores cancerous cells that arise in the body, as long as these cells continue to express self structures that the immune system has been trained to ignore. Dysfunctional tolerance is at the root of most autoimmune diseases, discussed further at the end of this chapter and in greater detail in Chapter 16. As one might imagine, failures in the establishment or maintenance of tolerance can have devastating clinical outcomes. The Immune Response Is Composed of Two Interconnected Arms: Innate Immunity and Adaptive Immunity Although reference is made to “the immune system,” it is important to appreciate that there are really two interconnected systems of immunity: innate and adaptive. These two systems collaborate to protect the body against foreign invaders. Innate immunity includes built-in molecular and cellular mechanisms that are encoded in the germline and are evolutionarily more primitive, aimed at preventing infection or quickly eliminating common invaders (Chapter 5). This includes physical and chemical barriers to infection, as well as the DNA-encoded receptors recognizing common chemical structures of many pathogens (see PRRs, above). In this case, rapid recognition and phagocytosis or destruction of the pathogen is the outcome. Innate immunity also includes a series of preexisting serum proteins, collectively referred to as complement, that bind common pathogenassociated structures and initiate a cascade of labeling and destruction events (Chapter 6). This highly effective first line of defense prevents most pathogens from taking hold, or eliminates infectious agents within hours of encounter. The recognition elements of the innate immune system are fast, some occurring within seconds of a barrier breach, but they are not very specific and are therefore unable to distinguish between small differences in foreign antigens. A second form of immunity, known as adaptive immunity, is much more attuned to subtle molecular differences. This part of the system, which relies on B and T lymphocytes, takes longer to come on board but is much more antigen specific. Typically, there is an adaptive immune response against a pathogen within 5 or 6 days after the barrier breach and initial exposure, followed by a gradual resolution of the infection. Adaptive immunity is slower partly because fewer cells possess the perfect receptor for the job: the antigenspecific, randomly generated receptors found on B and T cells. It is also slower because parts of the adaptive response rely on prior encounter and “categorizing” of antigens undertaken by innate processes. After antigen encounter, T and B lymphocytes undergo selection and proliferation, described earlier in the clonal selection theory of antigen specificity. Although slow to act, once these B and T cells have been selected and have honed their attack strategy, they become formidable opponents that can typically resolve the infection. The adaptive arm of the immune response evolves in real time in response to infection and adapts (thus the name) to better recognize, eliminate, and remember the invading pathogen. Adaptive responses involve a complex and interconnected system of cells and chemical signals that come together to finish the job initiated during the innate immune response. The goal of all vaccines against infectious disease is to elicit the development of specific and long-lived adaptive responses, so that the vaccinated individual will be protected in the future when the real pathogen comes along. This arm of immunity is orchestrated mainly via B and T lymphocytes following engagement of their randomly generated antigen recognition receptors. How these receptors are generated is a fascinating story, covered in detail in Chapter 7 of this book. An explanation of how these cells develop to maturity (Chapters 9 and 10) and then work in the body to protect us from infection (Chapters 11-14) or sometimes fail us (Chapters 15-19) takes up the vast majority of this book. The number of pages dedicated to discussing adaptive responses should not give the impression that this arm of the immune response is more important, or can work independently from, innate immunity. In fact, the full development of the adaptive response is dependent upon earlier innate pathways. The intricacies of their interconnections remain an area of intense study. The 2011 Nobel Prize in Physiology or Medicine was awarded to three scientists who helped clarify these two arms of the response: Bruce Beutler and Jules Hoffmann for discoveries related to the activation events important for innate immunity, and Ralph Steinman for his discovery of the role of dendritic cells in activating adaptive immune | responses (see Table 1-2). Because innate pathways make first contact with pathogens, the cells and molecules involved in this arm of the response use information gathered from their early encounter with pathogen to help direct the process of adaptive immune development. Adaptive immunity thus provides a second and more comprehensive line of defense, informed by the struggles undertaken by the innate system. It is worth noting that some infections are, in fact, eliminated by innate immune mechanisms alone, especially those that remain localized and involve very low numbers of fairly benign foreign invaders. (Think of all those insect bites or splinters in life that introduce bacteria under the skin!) Table 1-4 compares the major characteristics that distinguish innate and adaptive immunity. Although for ease of discussion the immune system is typically divided into these two arms of the response, there is considerable overlap of the cells and mechanisms involved in each of these arms of immunity. For innate and adaptive immunity to work together, these two systems must be able to communicate with one another. This communication is achieved by both cell-cell contact and by soluble messengers. Most of these soluble proteins are growth factor–like molecules known by the general name cytokines. Cytokines and cell surface ligands can bind with receptors found on responding cells and signal these cells to perform new functions, such as synthesis of other soluble factors or differentiation to a new cell type. A subset of these soluble signals are called chemokines because they have chemotactic activity, meaning they can recruit specific cells to the site. In this way, cytokines, chemokines, and other soluble factors produced by immune cells recruit or instruct cells and soluble proteins important for eradication of the pathogen from within the infection site. We’ve probably all felt this convergence in the form of swelling, heat, and tenderness at the site of exposure. These events are a part of a larger process collectively referred to as an inflammatory response, which is covered in detail in Chapter 15. Adaptive Immune Responses Typically Generate Memory One particularly significant and unique attribute of the adaptive arm of the immune response is immunologic TABLE 1-4 Magnitude of immune response Overview of the Immune System CHAPTER 1 17 Repeat Antigen A Antigen A Adaptive Innate 0 14 28 0 14 28 Time, days Primary response Secondary response FIGURE 1-8 Differences in the primary and secondary response to injected antigen reflect the phenomenon of immunologic memory. When an animal is injected with an antigen, it produces a primary antibody response (dark blue) of low magnitude and short duration, peaking at about 10 to 20 days. At some later point, a second exposure to the same antigen results in a secondary response that is greater in magnitude, peaks in less time (1–4 days), and is more antigen specific than the primary response. Innate responses, which have no memory element and occur each time an antigen is encountered, are unchanged regardless of how frequently this antigen has been encountered in the past (light blue). memory. This is the ability of the immune system to respond much more swiftly and with greater efficiency during a second exposure to the same pathogen. Unlike almost any other biological system, the vertebrate immune response has evolved not only the ability to learn from (adapt to) its encounters with foreign antigen in real time but also the ability to store this information for future use. During a first encounter with foreign antigen, adaptive immunity undergoes what is termed a primary response, during which the key lymphocytes that will be used to eradicate the pathogen are clonally selected, honed, and enlisted to resolve the infection. As mentioned above, these cells incorporate messages received from the innate players into their tailored response to the specific pathogen. All subsequent encounters with the same antigen or pathogen are referred to as the secondary response (Figure 1-8). Comparison of innate and adaptive immunity Innate Adaptive Response time Minutes to hours Days Specificity Limited and fixed Highly diverse; adapts to improve during the course of immune response Response to repeat infection Same each time More rapid and effective with each subsequent exposure Major components Barriers (e.g., skin); phagocytes; pattern recognition molecules T and B lymphocytes; antigen-specific receptors; antibodies 18 PA R T I | Introduction 1-9 OVERVIEW FIGURE Collaboration Between Innate and Adaptive Immunity in Resolving an Infection 2a 1 Lymph node P 3 P P B B T T 5a T B 4 5b Memory 2b This very basic scheme shows the sequence of events that occurs during an immune response, highlighting interactions between innate and adaptive immunity. 1. Pathogens are introduced at a mucosal surface or breach in skin (bacteria entering the throat, in this case), where they are picked up by phagocytic cells (yellow). 2a. In this innate stage of the response, the phagocytic cell undergoes changes and carries pieces of bacteria to a local lymph node to help activate adaptive immunity. 2b. Meanwhile, at the site of infection resident phagocytes encountering antigen release chemokines and cytokines (black dots) that cause fluid influx and help recruit other immune cells to the site (inflammation). 3. In the lymph node, T (blue) and B (green) cells with appropriate receptor specificity are clonally selected when During a secondary response, memory cells, kin of the final and most efficient B and T lymphocytes trained during the primary response, are re-enlisted to fight again. These cells begin almost immediately and pick up right where they left off, continuing to learn and improve their eradication strategy during each subsequent encounter with the same antigen. Depending on the antigen in question, memory cells can remain for decades after the conclusion of the primary response. Memory lymphocytes provide the means for subsequent responses that are so rapid, antigen-specific, and effective that when the same pathogen infects the body a second time, dispatch of the offending organism often occurs without symptoms. It is the remarkable property of memory that prevents us from catching many diseases a second time. Immunologic memory harbored by residual B and T lymphocytes is the foundation for vaccination, which uses crippled or killed pathogens as a safe way to “educate” the immune system to prepare it for later attacks by life-threatening pathogens. Overview Figure 1-9 highlights the ways in which the innate and adaptive immune responses work together to N B T their surface receptors bind antigen that has entered the system, kicking off adaptive immunity. 4. Collaboration between T and B cells and continued antigen encounter occurs in the lymph node, driving lymphocyte proliferation and differentiation, generating cells that can very specifically identify and eradicate the pathogen. For example: 5a. B cells secrete antibodies specific for the antigen, which travels to the site of infection to help label and eradicate the pathogen. 5b. In addition to the cells that will destroy the pathogen here, memory T and B cells are generated in this primary response and will be available at the initiation of a secondary response, which will be much more rapid and antigen specific. (Abbreviations: T  T cell, B  B cell, P = phagocyte; N = neutrophil, a type of immune cell.) resolve an infection. In this example, bacteria breach the mucosal lining of the throat, a skin or mucous barrier, where it is recognized and engulfed by a local phagocytic cell (step 1). As part of the innate immune system, the phagocytic cell releases cytokines and chemokines that attract other white blood cells to the site of infection, initiating inflammation (step 2b). That phagocytic cell may then travel to a local lymph node, the tissue where antigen and lymphocytes meet, carrying bacterial antigens to B and T lymphocytes (step 2a). Those lymphocytes with receptors that are specific for the antigen are selected, activated, and begin the adaptive immune response by proliferating (step 3). Activated TH cells help to activate B cells, and clonal expansion of both types of lymphocyte occurs in the lymph node (step 4). This results in many T and B cells specific for the antigen, with the latter releasing antibodies that can attach to the intruder and direct its destruction (step 5a). The adaptive response leaves behind memory T and B cells available for a future, secondary encounter with this antigen (step 5b). It is worth noting that memory is a unique Overview of the Immune System capacity that arises from adaptive responses; there is no memory component of innate immunity. Sometimes, as is the case for some vaccines, one round of antigen encounter and adaptation is not enough to impart protective immunity from the pathogen in question. In many of these cases, immunity can develop after a second or even a third round of exposure to an antigen. It is these sorts of pathogens that necessitate the use of vaccine booster shots. Booster shots are nothing more than a second or third episode of exposure to the antigen, each driving a new round of adaptive events (secondary response) and refinements in the responding lymphocyte population. The aim is to hone these responses to a sufficient level to afford protection against the real pathogen at some future date. The Good, Bad, and Ugly of the Immune System The picture we’ve presented so far depicts the immune response as a multicomponent interactive system that always protects the host from invasion by all sorts of pathogens. However, failures of this system do occur. They can be dramatic and often garner a great deal of attention, despite the fact that they are generally rare. Certain clinical situations also pose unique challenges to the immune system, including tissue transplants between individuals (probably not part of any evolutionary plan!) and the development of cancer. In this section we briefly describe some examples of common failures and challenges to the development of healthy immune responses. Each of these clinical manifestations is covered in much greater detail in the concluding chapters of this book (Chapters 15–19). Inappropriate or Dysfunctional Immune Responses Can Result in a Range of Disorders Most instances of immune dysfunction or failure fall into one of the following three broad categories: • Hypersensitivity (including allergy): overly zealous attacks on common benign but foreign antigens • Autoimmune Disease: erroneous targeting of selfproteins or tissues by immune cells • Immune Deficiency: insufficiency of the immune response to protect against infectious agents A brief overview of these situations and some examples of each are presented below. At its most basic level, immune dysfunction occurs as a result of improper regulation that allows the immune system to either attack something it shouldn’t or fail to attack something it should. Hypersensitivities, including allergy, and autoimmune disease are cases of the former, where the immune system attacks an improper | CHAPTER 1 19 target. As a result, the symptoms can manifest as pathological inflammation—an influx of immune cells and molecules that results in detrimental symptoms, including chronic inflammation and rampant tissue destruction. In contrast, immune deficiencies, caused by a failure to properly deploy the immune response, usually result in weakened or dysregulated immune responses that can allow pathogens to get the upper hand. This is a good time to mention that the healthy immune response involves a balancing act between immune aggression and immune suppression pathways. While we rarely fail to consider erroneous attacks (autoimmunity) or failures to engage (immune deficiency) as dysfunctional, we sometimes forget to consider the significance of the suppressive side of the immune response. Imperfections in the inhibitory arm of the immune response, present as a check to balance all the immune attacks we constantly initiate, can be equally profound. Healthy immune responses must therefore be viewed as a delicate balance, spending much of the time with one foot on the brakes and one on the gas. Hypersensitivity Reactions Allergies and asthma are examples of hypersensitivity reactions. These result from inappropriate and overly active immune responses to common innocuous environmental antigens, such as pollen, food, or animal dander. The possibility that certain substances induce increased sensitivity (hypersensitivity) rather than protection was recognized in about 1902 by Charles Richet, who attempted to immunize dogs against the toxins of a type of jellyfish. He and his colleague Paul Portier observed that dogs exposed to sublethal doses of the toxin reacted almost instantly, and fatally, to a later challenge with even minute amounts of the same toxin. Richet concluded that a successful vaccination typically results in phylaxis (protection), whereas anaphylaxis (anti-protection)—an extreme, rapid, and often lethal overreaction of the immune response to something it has encountered before—can result in certain cases in which exposure to antigen is repeated. Richet received the Nobel Prize in 1913 for his discovery of the anaphylactic response (see Table 1-2). The term is used today to describe a severe, life-threatening, allergic response. Fortunately, most hypersensitivity or allergic reactions in humans are not rapidly fatal. There are several different types of hypersensitivity reactions; some are caused by antibodies and others are the result of T-cell activity (see Chapter 15). However, most allergic or anaphylactic responses involve a type of antibody called immunoglobulin E (IgE). Binding of IgE to its specific antigen (allergen) induces the release of substances that cause irritation and inflammation, or the accumulation of cells and fluid at the site. When an allergic individual is exposed to an allergen, symptoms may include sneezing, wheezing (Figure 1-10), and difficulty in breathing (asthma); dermatitis or skin eruptions (hives); and, in more severe cases, strangulation due to constricted airways 20 PA R T I | Introduction BOX 1-3 CLINICAL FOCUS The Hygiene Hypothesis Worldwide, 300 million people suffer from asthma and approximately 250,000 people died from the disease in 2007 (see Chapter 15). As of 2009, in the United States alone, approximately 1 in 12 people (8.2%) are diagnosed with asthma. The most common reason for a trip to a hospital emergency room (ER) is an asthma attack, accounting for one-third of all visits. In addition to those treated in the ER, over 400,000 hospitalizations for asthma occurred in the United States in 2006, with an average stay of 3 to 4 days. In the past 25 years, the prevalence of asthma in industrialized nations has doubled. This is coupled with an overall rise in other types of allergic disease during the same time frame. What accounts for this climb in asthma and allergy in the last few decades? One idea, called the hygiene hypothesis, suggests that a decrease in human exposure to environmental microbes has had adverse effects on the human immune system. The hypothesis suggests that several categories of allergic or inflammatory disease, all disorders caused by excessive immune activation, have become more prevalent in industrialized nations thanks to diminished exposure to particular classes of microbes following the widespread use of antibiotics, immunization programs, and overall hygienic practices in those countries. This idea was first proposed by D. P. Strachan and colleagues in an article published in 1989 suggesting a link between hay fever and household hygiene. More recently, this hypothesis has been expanded to include the view by some that it may be a contributing factor in many allergic diseases, several autoimmune disorders, and, more recently, inflammatory bowel disease. What is the evidence supporting the hygiene hypothesis? The primary clinical support comes from studies that have shown a positive correlation between growing up under environmental conditions that favor microbe-rich (sometimes called “dirty”) environments and a decreased incidence of allergy, especially asthma. To date, childhood exposure to cowsheds and farm animals, having several older siblings, attending day care early in life, or growing up in a developing nation have all been correlated with a decreased likelihood of developing allergies later in life. While viral exposures during childhood do not seem to favor protection, exposure to certain classes of bacteria and parasitic organisms may. Of late, the primary focus of attention has been on specific classes of parasitic worms (called helminthes), spawning New Age allergy therapies involving intentional exposure. This gives whole new meaning to the phrase “Go eat worms”! What are the proposed immunologic mechanisms that might underlie this link between a lack of early-life microbial exposure and allergic disease? Current dogma supporting this hypothesis posits that millions of years of coevolution of microbes and humans has favored a system in which early exposure to a broad range of common environmental bugs helps set the immune system on a path of homeostatic balance between aggression and inhibition. Proponents of this immune regulation argument, sometimes referred to as the “old friends” hypothesis, suggest that antigens present on microbial organisms that have played a longstanding role in our evolutionary history (both pathogens and harmless microbes we ingest or that make up our historical flora) may engage with the pattern recognition receptors (PPRs) present on cells of our innate immune system, driving them to warn cells involved in adaptive responses to tone it down. This hypothesis posits that without early and regular exposure of our immune cells to these old friends and their antigens, the development of “normal” immune regulatory or homeostatic responses is thrown into disarray, setting us up for an immune system poised to overreact in the future. Animal models of disease lend some support to this hypothesis and have helped immunologists probe this line of thinking. For instance, certain animals raised in partially or totally pathogen-free environments are more prone to type 1, or insulin-dependent, diabetes, an autoimmune disease caused by immune attack of pancreatic cells (see Chapter 16). The lower the infectious burden of exposure in these mice, the greater the incidence of diabetes. Animals specifically bred to carry enhanced genetic susceptibility favoring spontaneous development of diabetes (called NOD mice, for nonobese diabetic) and treated with a variety of infectious agents can be protected from diabetes. Meanwhile, NOD mice maintained in pathogen-free housing almost uniformly develop diabetes. Much like this experimental model, susceptibility to asthma and most other allergies is known to run in families. Although all the genes linked to asthma have not yet been characterized, it is known that you have a 30% chance of developing the disease if one of your parents is a sufferer, and a 70% chance if both parents have asthma. While the jury may still be out concerning the verdict behind the hygiene hypothesis, animal and human studies clearly point to strong roles for both genes and environment in susceptibility to allergy. As data in support of this hypothesis continue to grow, the old saying concerning a dirty child—that “It’s good for their immune system”—may actually hold true! Centers for Disease Control and Prevention. 2012. CDC: Preventing Chronic Disease 9: 110054. Liu, A. H., and Murphy, J. R. 2003. Hygiene hypothesis: Fact or fiction? Journal of Allergy and Clinical Immunology 111(3):471–478. Okada, H., et al. 2010. The hygiene hypothesis for autoimmune and allergic diseases: An update. Clinical and Experimental Immunology 160:1. Sironi, M., and M. Clerici. 2010. The hygiene hypothesis: An evolutionary perspective. Microbes and Infection 12:421. Overview of the Immune System FIGURE 1-10 Patient suffering from hay fever as a result of an allergic reaction. Such hypersensitivity reactions result from sensitization caused by previous exposure to an antigen in some individuals. In the allergic individual, histamines are released as a part of the hypersensitivity response and cause sneezing, runny nose, watery eyes, and such during each subsequent exposure to the antigen (now called an allergen) [Source: Chris Rout/Alamy.] following extreme inflammation. A significant fraction of our health resources is expended to care for those suffering from allergies and asthma. One particularly interesting rationale to explain the unexpected rise in allergic disease is called the hygiene hypothesis and is discussed in the Clinical Focus on page 20. Autoimmune Disease Sometimes the immune system malfunctions and a breakdown in self-tolerance occurs. This could be caused by a sudden inability to distinguish between self and nonself or by a misinterpretation of a self-component as dangerous, causing an immune attack on host tissues. This condition, called autoimmunity, can result in a number of chronic debilitating diseases. The symptoms of autoimmunity differ, depending on which tissues or organs are under attack. For example, multiple sclerosis is due to an autoimmune attack on a protein in nerve sheaths in the brain and central nervous system that results in neuromuscular dysfunction. Crohn’s disease is an attack on intestinal tissues that leads to destruction of gut epithelia and poor absorption of food. One of the most common autoimmune disorders, rheumatoid arthritis, results from an immune attack on joints of the hands, feet, arms, and legs. Both genetic and environmental factors are likely involved in the development of most autoimmune diseases. However, the exact combination of genes and environmental exposures that favor the development of a particular autoimmune disease are difficult to pin down, and constitute very active areas of immunologic research. Recent discoveries in these areas and the search for improved treatments are all covered in greater detail in Chapter 16. | CHAPTER 1 21 Immune Deficiency In most cases, when a component of innate or adaptive immunity is absent or defective, the host suffers from some form of immunodeficiency. Some of these deficiencies produce major clinical effects, including death, while others are more minor or even difficult to detect. Immune deficiency can arise due to inherited genetic factors (called primary immunodeficiencies) or as a result of disruption/damage by chemical, physical, or biological agents (termed secondary immunodeficiencies). Both of these forms of immune deficiency are discussed in greater detail in Chapter 18. The severity of the disease resulting from immune deficiency depends on the number and type of affected immune response components. A common type of primary immunodeficiency in North America is a selective immunodeficiency in which only one type of antibody, called Immunoglobulin A is lacking; the symptoms may be an increase in certain types of infections, or the deficiency may even go unnoticed. In contrast, a more rare but much more extreme deficiency, called severe combined immunodeficiency (SCID), affects both B and T cells and basically wipes out adaptive immunity. When untreated, SCID frequently results in death from infection at an early age. By far, the most common form of secondary immunodeficiency is Acquired Immune Deficiency Syndrome (AIDS), resulting from infection with Human Immunodeficiency Virus (HIV). As discussed further in Chapter 18, humans do not effectively recognize and eradicate this virus. Instead, a state of persistent infection occurs, with HIV hiding inside the genomes of TH cells, its target cell type and the immune cell type that is critical to guiding the direction of the adaptive immune response. As the immune attack on the virus mounts, more and more of these TH cells are lost. When the disease progresses to AIDS, so many TH cells have been destroyed or otherwise rendered dysfunctional that a gradual collapse of the immune system occurs. It is estimated that at the end of 2010, more than 34 million people worldwide suffered from this disease (for more current numbers, see www.unaids.org), which if not treated can be fatal. For patients with access, certain anti-retroviral treatments can now prolong life with HIV almost indefinitely. However, there is neither a vaccine nor a cure for this disease. It is important to note that many pervasive pathogens in our environment cause no problem for healthy individuals thanks to the immunity that develops following initial exposure. However, individuals with primary or secondary deficiencies in immune function become highly susceptible to disease caused by these ubiquitous microbes. For example, the fungus Candida albicans, present nearly everywhere and a nonissue for most individuals, can cause an irritating rash and a spreading infection in the mucosal surface of the mouth and vagina in patients suffering from immune deficiency (see Figure 1-6b). The resulting rash, called thrush, can sometimes be the first sign of immune dysfunction (Figure 1-11). If left unchecked, C. albicans can spread, causing systemic candidiasis, a life-threatening condition. Such 22 PA R T I | Introduction FIGURE 1-11 An immune deficient patient suffering from oral thrush due to opportunistic infection by Candida albicans. [Creative Commons <http://en.wikipedia.org/wiki/File: Thrush.JPG] infections by ubiquitous microorganisms that cause no harm in an immune competent host but which are often observed only in cases of underlying immune deficiency, are termed opportunistic infections. Several rarely seen opportunistic infections identified in early AIDS patients were the first signs that these patients had seriously compromised immune systems, and helped scientists to identify the underlying cause. Unchecked opportunistic infections are still the main cause of death in AIDS patients. The Immune Response Renders Tissue Transplantation Challenging Normally, when the immune system encounters foreign cells, it responds strongly to rid the host of the presumed invader. However, in the case of transplantation, these cells or tissues from a donor may be the only possible treatment for lifethreatening disease. For example, it is estimated that more than 70,000 persons in the United States alone would benefit from a kidney transplant. The fact that the immune system will attack and reject any transplanted organ that is nonself, or not a genetic match, raises a formidable barrier to this potentially lifesaving treatment, presenting a particularly unique challenge to clinicians who treat these patients. While the rejection of this transplant by the recipient’s immune system may be seen as a “failure,” in fact it is just a consequence of the immune system functioning properly. Normal tolerance processes governing self/nonself discrimination and immune engagement caused by danger signals (partially the result of the trauma caused by surgical transplantation) lead to the rapid influx of immune cells and coordinated attacks on the new resident cells. Some of these transplant rejection responses can be suppressed using immune inhibitory drugs, but treatment with these drugs also suppresses general immune function, leaving the host susceptible to opportunistic infections. Research related to transplantation studies has played a major role in the development of the field of Immunology. A Nobel Prize was awarded to Karl Landsteiner (mentioned earlier for his contributions to the concept of immune specificity) in 1930 for the discovery of the human ABO blood groups, a finding that allowed blood transfusions to be carried out safely. In 1980, G. Snell, J. Dausset, and B. Benacerraf were recognized for discovery of the major histocompatibility complex (MHC). These are the tissue antigens that differ most between non-genetically identical individuals, and are thus one of the primary targets of immune rejection of transplanted tissues. Finally, in 1991 E. D. Thomas and J. Murray were awarded Nobel prizes for treatment advances that paved the way for more clinically successful tissue transplants (see Table 1-2). Development of procedures that would allow a foreign organ or cells to be accepted without suppressing immunity to all antigens still remains a major goal, and a challenge, for immunologists today (see Chapter 16). Cancer Presents a Unique Challenge to the Immune Response Cancer, or malignancy, occurs in host cells when they begin to divide out of control. Since these cells are self in origin, self-tolerance mechanisms can inhibit the development of an immune response, making the detection and eradication of cancerous cells a continual challenge. That said, it is clear that many tumor cells do express unique or developmentally inappropriate proteins, making them potential targets for immune cell recognition and elimination, as well as targets for therapeutic intervention. However, as with many microbial pathogens, the increased genetic instability of these rapidly dividing cells gives them an advantage in terms of evading immune detection and elimination machinery. We now know that the immune system actively participates in the detection and control of cancer in the body (see Chapter 19). The number of malignant disorders that arise in individuals with compromised immunity highlights the degree to which the immune system normally controls the development of cancer. Both innate and adaptive elements have been shown to be involved in this process, although adaptive immunity likely plays a more significant role. However, associations between inflammation and the development of cancer, as well as the degree to which cancerous cells evolve to become more aggressive and evasive under pressure from the immune system, have demonstrated that the immune response to cancer can have both healing and disease-inducing characteristics. As the mechanics of these elements are resolved in greater detail, there is hope that therapies can be designed to boost or maximize the anti-tumor effects of Overview of the Immune System immune cells while dampening their tumor enhancing activities. Our understanding of the immune system has clearly come a very long way in a fairly short time. Yet much still remains to be learned about the mammalian immune response and the ways in which this system interacts with other body systems. With this enhanced knowledge, the | CHAPTER 1 23 hope is that we will be better poised to design ways to modulate these immune pathways through intervention. This would allow us to develop more effective prevention and treatment strategies for cancer and other diseases that plague society today, not to mention preparing us to respond quickly to the new diseases or infectious agents that will undoubtedly arise in the future. S U M M A R Y ■ ■ ■ ■ ■ ■ ■ ■ ■ Immunity is the state of protection against foreign pathogens or substances (antigens). Vaccination is a means to prepare the immune system to effectively eradicate an infectious agent before it can cause disease, and its widespread use has saved many lives. Humoral immunity involves combating pathogens via antibodies, which are produced by B cells and can be found in bodily fluids. Antibodies can be transferred between individuals to provide passive immune protection. Cell-mediated immunity involves primarily antigenspecific T lymphocytes, which act to eradicate pathogens or otherwise aid other cells in inducing immunity. Pathogens fall into four major categories and come in many forms. The immune response quickly becomes tailored to the type of organism involved. The immune response relies on recognition molecules that can be germline encoded or randomly generated. The process of self-tolerance ensures that the immune system avoids destroying host tissue. The vertebrate immune response can be divided into two interconnected arms of immunity: innate and adaptive. Innate responses are the first line of defense, utilizing germline-encoded recognition molecules and phagocytic ■ ■ ■ ■ ■ ■ cells. Innate immunity is faster but less specific than adaptive responses, which take several days but are highly antigen specific. Innate and adaptive immunity operate cooperatively; activation of the innate immune response produces signals that stimulate and direct subsequent adaptive immune pathways. Adaptive immunity relies upon surface receptors, called B- and T-cell receptors, that are randomly generated by DNA rearrangements in developing B and T cells. Clonal selection is the process by which individual T and B lymphocytes are engaged by antigen and cloned to create a population of antigen-reactive cells. Memory cells are residual B and T cells that remain after antigen exposure and that pick up where they left off during a subsequent, or secondary, response. Dysfunctions of the immune system include common maladies such as allergies, asthma, and autoimmune disease (overly active or misdirected immune responses) as well as immune deficiency (insufficient immune responses). Transplanted tissues and cancer present unique challenges to clinicians, because the healthy immune system typically rejects or destroys nonself proteins, such as those encountered in most transplant situations, and tolerates self cells. R E F E R E N C E S Burnet, F. M. 1959. The Clonal Selection Theory of Acquired Immunity. Cambridge University Press, Cambridge, England. Paul, W., ed. 2003. Fundamental Immunology, 5th ed. Lippincott Williams & Wilkins, Philadelphia. Desour, L. 1922. Pasteur and His Work (translated by A. F. and B. H. Wedd). T. Fisher Unwin, London. Silverstein, A. M. 1979. History of immunology. Cellular versus humoral immunity: determinants and consequences of an epic 19th century battle. Cellular immunology. 48:208. Kimbrell, D. A., and B. Beutler. 2001. The evolution and genetics of innate immunity. Nature Reviews Genetics 2:256. Kindt, T. J., and J. D. Capra. 1984. The Antibody Enigma. Plenum Press, New York. Landsteiner, K. 1947. The Specificity of Serologic Reactions. Harvard University Press, Cambridge, MA. Medawar, P. B. 1958. The Immunology of Transplantation: The Harvey Lectures, 1956–1957. Academic Press, New York. Metchnikoff, E. 1905. Immunity in the Infectious Diseases. Macmillan, New York. Useful Web Sites www.aai.org The Web site of the American Association of Immunologists contains a good deal of information of interest to immunologists. www.ncbi.nlm.nih.gov/PubMed PubMed, the National Library of Medicine database of more than 9 million publications, is the world’s most comprehensive bibliographic 24 PA R T I | Introduction database for biological and biomedical literature. It is also a highly user-friendly site. www.nobelprize.org/nobel_prizes/medicine/ laureates The official Web site of the Nobel Prize in Physi- www.aaaai.org The American Academy of Allergy ology or Medicine. Asthma and Immunology site includes an extensive library of information about allergic diseases. www.historyofvaccines.org A Web site run by The www.who.int/en The World Health Organization directs and coordinates health-related initiatives and collects worldwide health statistics data on behalf of the United Nations system. www.cdc.gov Part of the United States Department of Health and Human Services, the Centers for Disease Control and Prevention coordinates health efforts in the United States and provides statistics on U.S. health and disease. S T U D Y College of Physicians of Philadelphia with facts, articles, and timelines related to vaccine developments. www.niaid.nih.gov The National Institute of Allergy and Infectious Disease is a branch of the U.S. National Institute of Health that specifically deals with research, funding, and statistics related to basic immunology, allergy, and infectious disease threats. Q U E S T I O N S 1. Why was Jenner’s vaccine superior to previous methods for e. Innate immunity is deployed only during the primary conferring resistance to smallpox? 2. Did the treatment for rabies used by Pasteur confer active or passive immunity to the rabies virus? Is there any way to test this? f. g. 3. Infants immediately after birth are often at risk for infec- tion with group B Streptococcus. A vaccine is proposed for administration to women of childbearing years. How can immunizing the mothers help the babies? 4. Indicate to which branch(es) of the immune system the following statements apply, using H for the humoral branch and CM for the cell-mediated branch. Some statements may apply to both branches (B). a. Involves B cells b. Involves T cells c. Responds to extracellular bacterial infection d. Involves secreted antibody e. Kills virus-infected self cells 5. Adaptive immunity exhibits several characteristic attributes, which are mediated by lymphocytes. List four attributes of adaptive immunity and briefly explain how they arise. 6. Name three features of a secondary immune response that distinguish it from a primary immune response. 7. Give examples of mild and severe consequences of immune dysfunction. What is the most common cause of immunodeficiency throughout the world today? 8. For each of the following statements, indicate whether the statement is true or false. If you think the statement is false, explain why. a. Booster shots are required because repeated exposure to an antigen builds a stronger immune response. b. The gene for the T cell receptor must be cut and spliced together before it can be expressed. c. Our bodies face the greatest onslaught from foreign invaders through our skin. d. Increased production of antibody in the immune system is driven by the presence of antigen. h. i. j. response, and adaptive immunity begins during a secondary response. Autoimmunity and immunodeficiency are two different terms for the same set of general disorders. If you receive intravenous immunoglobulin to treat a snakebite, you will be protected from the venom of this snake in the future, but not venom from other types of snakes. Innate and adaptive immunity work collaboratively to mount an immune response against pathogens. The genomic sequences in our circulating T cells for encoding a T-cell receptor are the same as those our parents carry in their T cells. Both the innate and adaptive arms of the immune response will be capable of responding more efficiently during a secondary response. 9. What was the significance of the accidental re-inoculation of some chickens that Pasteur had previously exposed to the bacteria that causes cholera? Why do you think these chickens did not die after the first exposure to this bacterium? 10. Briefly describe the four major categories of pathogen. Which are likely to be the most homogenous in form and which the most diverse? Why? 11. Describe how the principle of herd immunity works to protect unvaccinated individuals. What characteristics of the pathogen or of the host do you think would most impact the degree to which this principle begins to take hold? 12. What is the difference between the discarded instructional theory for lymphocyte specificity and the selection theory, which is now the accepted explanation? 13. Compare and contrast innate and adaptive immunity by matching the following characteristics with the correct arm of immunity, using I for innate and A for adaptive: a. Is the first to engage upon initial encounter with antigen b. Is the most pathogen specific Overview of the Immune System c. Employs T and B lymphocytes d. Adapts during the response e. Responds identically during a first and second exposure f. g. h. i. j. k. to the same antigen Responds more effectively during a subsequent exposure Includes a memory component Is the target of vaccination Can involve the use of PAMP receptors Involves antigen-specific receptors binding to pathogens Can be mediated by antibodies | CHAPTER 1 25 14. What is meant by the term tolerance? How do we become tolerant to the structures in our own bodies? 15. What is an antigen? An antibody? What is their relation- ship to one another? 16. How are PRRs different from B- or T-cell receptors? Which is most likely to be involved in innate immunity and which in adaptive immunity? 17. In general terms, what role do cytokines play in the develop- ment of immunity? How does this compare to chemokines? This page lelt intentionally blank. 2 Cells, Organs, and Microenvironments of the Immune System A successful immune response to a pathogen depends on finely choreographed interactions among diverse cell types (see Figure 1-9): innate immune cells that mount a first line of defense against pathogen, antigen-presenting cells that communicate the infection to lymphoid cells, which coordinate the adaptive response and generate the memory cells, which prevent future infections. The coordination required for the development of a full immune response is made possible by the specialized anatomy and microanatomy of the immune system, which is dispersed throughout the body and organizes cells in time and space. Primary lymphoid organs— including the bone marrow and the thymus—regulate the development of immune cells from immature precursors. Secondary lymphoid organs—including the spleen, lymph nodes, and specialized sites in the gut and other mucosal tissues—coordinate the encounter of antigen with antigen-specific lymphocytes and their development into effector and memory cells. Blood vessels and lymphatic systems connect these organs, uniting them into a functional whole. Remarkably, all functionally specialized, mature blood cells (red blood cells, granulocytes, macrophages, dendritic cells, and lymphocytes) arise from a single cell type, the hematopoietic stem cell (HSC) (Figure 2-1). The process by which HSCs differentiate into mature blood cells is called hematopoiesis. Two primary lymphoid organs are responsible for the development of stem cells into mature immune cells: the bone marrow, where HSCs reside and give rise to all cell types; and the thymus, where T cells complete their maturation. We will begin this chapter by describing the structure and function of each cell type that arises from HSCs, and the structure and function of both the bone marrow and thymus in the context of hematopoiesis and thymopoiesis. We will then describe the secondary lymphoid organs, where the immune response is initiated. The lymph nodes and spleen will be featured, but lymphoid tissue associated with mucosal layers will also be discussed. Four focused discussions are also Scanning electron micrograph of blood vessels in a lymph node. Susumu Nishinaga/ Photo Researchers ■ ■ ■ Cells of the Immune System Primary Lymphoid Organs—Where Immune Cells Develop Secondary Lymphoid Organs—Where the Immune Response Is Initiated included in this chapter. Specifically, in two Classic Experiment Boxes, we describe the discovery of a second thymus and the history behind the identification of hematopoietic stem cells. In a Clinical Focus Box, we discuss the clinical use and promise of hematopoietic stem cells, and finally, in an Evolution Box, we describe some intriguing variations in the anatomy of the immune system among our vertebrate relatives. Cells of the Immune System Stem cells are defined by two capacities: (1) the ability to regenerate or “self-renew” and (2) the ability to differentiate into all diverse cell types. Embryonic stem cells have the capacity to generate every specialized cell type in an organism (in other words, they are pluripotent). Adult stem cells, in contrast, have the capacity to give rise to the diverse cell types that specify a particular tissue. Multiple adult organs harbor stem cells (“adult stem cells”) that can give rise to mature tissue-specific cells. The HSC is considered the paradigmatic adult stem cell because it can differentiate into all the types of blood cells. 27 28 PA R T I | Introduction Peripheral tissues Bone-marrow Hematopoietic stem cell Peripheral tissues Selfrenewing Dendritic cell Myeloid progenitor Lymphoid progenitor Natural killer (NK) cell Thymus Macrophage Monocyte TH helper cell Neutrophil Granulocytemonocyte progenitor T-cell progenitor TC cytotoxic cell Eosinophil Eosinophil progenitor B-cell progenitor Mast cell Basophil B cell Basophil progenitor Dendritic cell Platelets Megakaryocyte Erythrocyte Erythroid progenitor FIGURE 2-1 Hematopoiesis. Self-renewing hematopoietic stem cells give rise to lymphoid and myeloid progenitors. Most immune cells mature in the bone marrow and then travel to peripheral organs via the blood. Some, including mast cells and macrophages, undergo further maturation outside the bone marrow. T cells develop to maturity in the thymus. Hematopoietic Stem Cells Have the Ability to Differentiate into Many Types of Blood Cells HSCs are rare—fewer than one HSC is present per 5  104 cells in the bone marrow—and their numbers are strictly controlled by a balance of cell division, death, and differentiation. Under conditions where the immune system is not being challenged by a pathogen (steady state or homeostatic conditions), most HSCs are quiescent. A small number divide, generating daughter cells. Some daughter cells retain the stem-cell characteristics of the mother cell—that is, they remain self-renewing and able to give rise to all blood cell types. Other daughter cells differentiate into progenitor cells that lose their self-renewal capacity and become progressively more committed to a particular blood cell lineage. As an organism ages, the number of HSCs decreases, demonstrating that there are limits to an HSC’s self-renewal potential. When there is an increased demand for hematopoiesis (e.g., during an infection or after chemotherapy), HSCs display an enormous proliferative capacity. This can be demonstrated in Cells, Organs, and Microenvironments of the Immune System mice whose hematopoietic systems have been completely destroyed by a lethal dose of x-rays (950 rads). Such irradiated mice die within 10 days unless they are infused with normal bone marrow cells from a genetically identical mouse. Although a normal mouse has 3  108 bone marrow cells, infusion of only 104 to 105 bone marrow cells from a donor is sufficient to completely restore the hematopoietic system, which demonstrates the enormous capacity of HSCs for self-renewal. Our ability to identify and purify this tiny subpopulation has improved considerably, and investigators can now theoretically rescue irradiated animals with just a few purified stem cells, which give rise to progenitors that proliferate rapidly and populate the blood system relatively quickly. Because of the rarity of HSCs and the challenges of culturing them in vitro, investigators initially found it very difficult to identify and isolate HSCs. The Classic Experiment Box on | CHAPTER 2 29 pages 29–31 describes experimental approaches that led to successful enrichment of HSCs. Briefly, the first successful efforts featured clever process-of-elimination strategies. Investigators reasoned that undifferentiated hematopoietic stem cells would not express surface markers specific for mature cells from the multiple blood lineages (“Lin” markers). They used several approaches to eliminate cells in the bone marrow that did express these markers (Lin cells) and then examined the remaining (Lin) population for its potential to continually give rise to all blood cells over the long term. Lin cells were, indeed, enriched for this potential. Other investigators took advantage of two technological developments that revolutionized immunological research—monoclonal antibodies and flow cytometry (see Chapter 20)—and identified surface proteins, including CD34, Sca-1, and c-Kit, that distinguished the rare hematopoietic stem cell population. BOX 2-1 CLASSIC EXPERIMENT Isolating Hematopoietic Stem Cells By the 1960s researchers knew that hematopoietic stem cells (HSCs) existed and were a rare population in the bone marrow. However, they did not know what distinguished them from the other millions of cells that crowded the bone marrow. It was clear that in order to fully understand how these remarkable cells give rise to all other blood cells, and in order to harness this potential for clinical use, investigators would have to find a way to isolate HSCs. But how do you find something that is very rare, whose only distinctive feature is its function—its ability to give rise to all blood cells? Investigators devised a variety of strategies that evolved rapidly with every technological advance, particularly with the advent of monoclonal antibodies and flow cytometry. Regardless of what method one uses to try to isolate a cell, it is critically important to have a reliable experimental assay that can tell you that the cells that you have teased out are, indeed, the ones you are looking for. Fundamentally, in order to prove that you have enriched or purified an HSC, you have to show that it can proliferate and give rise to all blood cell types in an animal over the long term. Many of the assays that were originally established to show this are still in use. They include colony formation assays, where the ability of individual cells to proliferate (and differentiate) is determined by looking for evidence of cell division either in vitro (on plates) or in vivo (in the spleens of irradiated mice). However, the best evidence for successful isolation of HSCs is the demonstration that they can restore the blood cells and immune system of a lethally irradiated animal, preventing its death. This can be done for mouse stem cells by injecting stem cell candidates into irradiated mice and determining if they confer survival and repopulate all blood cell types. The development of a mouse model that accepts human hematopoietic stem cells (the SCID-hu (man) mouse model) has greatly enhanced investigators’ ability to verify the pluripotentiality of candidate human stem cell populations. In the 1970s investigators did not have the ability to easily compare differences in protein and gene expression among single cells, so they had to try to distinguish cell types based on other physical and structural features. It wasn’t until monoclonal antibodies were introduced into research repertoires that investigators could seriously consider purifying a stem cell. Monoclonal antibodies (described in Chapter 20) can be raised to virtually any protein, lipid, or carbohydrate. Monoclo- nal antibodies can, themselves, be covalently modified with gold particles, enzymes, or fluorochromes in order to visualize their binding by microscopy. In the early 1980s, investigators reasoned that HSCs were unlikely to express proteins specific for mature blood cells. Using monoclonal antibodies raised against multiple mature cells, they trapped and removed them from bone marrow cell suspensions, first via a process called panning, where the heterogeneous pool of cells was incubated with antibodies bound to a plastic and then those cells that did not stick were dislodged and poured off. The cells that did not stick to the antibodies were, indeed, enriched (in some cases by several thousandfold) in cells with HSC potential. This negative selection strategy against mature cell lineages continues to be useful and is now referred to as “lineage or Lin” selection; cells enriched by this method are referred to as “Lin.” The panning process that yielded one of the first images of cells that included human hematopoietic stem cells is shown in Figure 1. Investigators also worked to identify surface molecules that were specific to hematopoietic stem cells, so that they could positively select them from the diverse bone marrow cell types. The first protein that identified human HSCs, now known as (continued) PA R T I 30 | CLASSIC EXPERIMENT Introduction (continued) B P E N S P L Eo P E M L B E P Add bone marrow suspension Coat plate with lineage specific antibodies Culture for 1-2 h S P E Eo N P L P B M P Discard what does bind E Eo N L Keep what doesn’t bind B P P S Lin– cells Stain and observe under microscope M FIGURE 1 Panning for stem cells. Early approaches to isolate HSCs took advantage of antibodies that were raised against mature blood cells and a process called panning, where cells are incubated in plastic plates that are coated with antibodies. Specifically, investigators layered a suspension of bone marrow onto plastic plates coated with antibodies that could bind multiple different mature (“lineage positive”) blood cells. They waited 1 to 2 hours and then washed the cells that did not stick (the immature, “lineage negative” blood cells) off the plate. Most cells—mature blood cells—did stick firmly to the plate. However, the cells that did not stick were enriched for hematopoietic stem cells. This process led to the first image of human bone marrow cells enriched for hematopoietic stem cells by panning. S = stem cell; P = progenitor cell; M  monocyte; B  basophil; N  neutrophil; Eo  eosinophil; L  lymphocyte; E  erythrocyte. [Emerson, S.G., Colin, A.S., Wang, E.A., Wong, G.G., Clark, S.C., and Nathan, D.G. Purification and Demonstration of Fetal Hematopoietic Progenitors and Demonstration of Recombinant Multipotential Colony-stimulating Activity. J. Clin. Invest., Vol. 76, Sept. 1985, 1286-1290. ©The American Society for Clinical Investigation.] CD34, was identified with a monoclonal antibody raised against a tumor of primitive blood cell types (acute myeloid leukemia). Although one can positively select cells from a diverse population using the panning procedures described above (or by using its more current variant where cells are applied to columns of resin-bound monoclonal antibodies or equivalents), the flow cytometer provides the most efficient way to pull out a rare population from a diverse group of cells. This machine, invented by the Herzenberg laboratory and its interdisciplinary team of scholars and inventors, has revolutionized immunology and clinical medicine. In a nutshell, it is a machine that allows one to identify, separate, and recover individual cells from a diverse pool of cells on the basis of the profiles of proteins and/or genes they express. Chapter 20 provides you with an under-the- hood introduction to the remarkable technology. Briefly, cells from a heterogeneous suspension are tagged with (“stained with”) monoclonal antibodies (or other molecules) that bind to distinct features and are coupled to distinct fluorochromes. These cells flow single file in front of lasers that excite the several fluorochromes, and the intensities of the multiple wavelengths given off by each individual cell are recorded. Cells that express specific antigens at desired levels (e.g., showing evidence for expression of CD34) can be physically separated from other cells and recovered for further studies. In the late 1980s, Irv Weissman and his laboratories discovered that differences in expression of the Thy protein (a T-cell marker), and later the expression of Sca protein, differentiated mouse hematopoietic stem cells from more mature cells. His laboratory combined both negative and positive selection techniques to develop one of the most efficient approaches for hematopoietic stem cell enrichment (Figure 2). As other surface molecules were identified, the approach was honed. Currently, HSCs are most frequently identified by their Lin Sca1 c-Kit (“LSK”) phenotype. It has become clear that even this subgroup, which represents less than 1% of bone marrow cells, is phenotypically and functionally heterogeneous. Other surface markers, including SLAM proteins, that can distinguish among these subpopulations continue to be identified. The synergy between technical developments and experimental strategies will undoubtedly continue so that, ultimately, we will be able to unambiguously identify, isolate, and manipulate what remains the holy grail of HSC investigations: the longterm stem cell that can both self-renew and give rise to all blood cell types. | Cells, Organs, and Microenvironments of the Immune System CHAPTER 2 31 BOX 2-1 (a) Whole bone marrow Lethally irradiated mouse (950 rads) 2 × 105 unenriched cells L P Eo L P E S B N E N M P React with Fl-antibodies to differentiation antigens Restore hematopoiesis, mouse lives Negative selection against cells that express mature (lineage) markers Lin- cells 1 × 103 partly enriched cells S M P P N E P L B E React with Fl-antibodies against Sca-1 and c-Kit Restore hematopoiesis, mouse lives Lin-Sca-1+c-Kit+ (LSK) cells L Eo Differentiated cells N Positive selection for cells that express stem cell markers (e.g., Sca-1 and c-Kit) 30–100 fully enriched cells P P Stem Progenitor cell cells Restore hematopoiesis, mouse lives (b) P 100 Survival rate, % S (Lin-Sca-1+c-Kit+ (LSK) cells Fully enriched cells Partly enriched cells (Lin–) Unenriched cells (Whole bone marrow) 101 102 103 104 105 Number of cells injected into lethally irradiated mouse FIGURE 2 Current approaches for enrichment of the pluripotent stem cells from bone marrow. A schematic of the type of stem cell enrichment now routinely employed, but originated by Irv Weissman and colleagues. (a) Enrichment is accomplished by (1) removal (negative selection) of differentiated hematopoietic cells (white) from whole bone marrow after treatment with fluorescently labeled antibodies (Fl-antibodies) specific for membrane molecules expressed on differentiated (mature) lineages but absent from the undifferentiated stem cells (S) and progenitor cells (P), followed by (2) retention (positive selection) of cells within the resulting partly enriched preparation that bound to antibodies specific for Sca-1 and c-Kit, two early differentiation antigens. Cells that are enriched by removal of differentiated cells are referred to as “lineage-minus” or Lin populations. Cells that are further enriched by positive selection according to Sca-1 and c-Kit expression are referred to as LSK (for LinSca-1c-Kit) cells. M  monocyte; B  basophil; N  neutrophil; Eo  eosinophil; L  lymphocyte; E  erythrocyte. (b) Enrichment of stem cell preparations is measured by their ability to restore hematopoiesis in lethally irradiated mice. Only animals in which hematopoiesis occurs survive. Progressive enrichment of stem cells (from whole bone marrow, to Lin populations, to LSK populations) is indicated by the decrease in the number of injected cells needed to restore hematopoiesis. A total enrichment of about 1000-fold is possible by this procedure. Emerson, S. G., C. A. Sieff, E. A. Wang, G. G. Wong, S. C. Clark, and D. G. Nathan. 1985. Purification of fetal hematopoietic progenitors and demonstration of recombinant multipotential colony-stimulating activity. Journal of Clinical Investigation 76:1286. Shizuru, J. A., R. S. Negrin, and I. L. Weissman. 2005. Hematopoietic stem and progenitor cells: clinical and preclinical regeneration of the hematolymphoid system. Annual Review of Medicine 56:509. Weissman, I. L. 2000. Translating stem and progenitor cell biology to the clinic: Barriers and opportunities. Science 287:1442. PA R T I 32 | Introduction Hematopoiesis Is the Process by Which Hematopoietic Stem Cells Develop into Mature Blood Cells An HSC that is induced to differentiate (undergo hematopoiesis) loses its self-renewal capacity and makes one of two broad lineage commitment choices (see Figure 2-1). It can become a common myeloid-erythroid progenitor (CMP), which gives rise to all red blood cells (the erythroid lineage), granulocytes, monocytes, and macrophages (the myeloid lineage), or it can become a common lymphoid progenitor (CLP), which gives rise to B lymphocytes, T lymphocytes, and NK cells. Myeloid cells and NK cells are members of the innate immune system, and are the first cells to respond to infection or other insults. Lymphocytes are members of the adaptive immune response and generate a refined antigenspecific immune response that also gives rise to immune memory. As HSCs progress along their chosen lineages, they lose the capacity to contribute to other cellular lineages. Interestingly, both myeloid and lymphoid lineages give rise to dendritic cells, antigen-presenting cells with diverse features and functions that play an important role in initiating adaptive immune responses. The concentration and frequency of immune cells in blood are listed in Table 2-1. Regulation of Lineage Commitment during Hematopoiesis Each step a hematopoietic stem cell takes toward commitment to a particular cellular lineage is accompanied by genetic changes. Multiple genes that specify lineage commitment have been identified. Many of these are transcriptional regulators. For instance, the transcription factor GATA-2 is required for the development of all hematopoietic lineages; in its absence animals die during embryogenesis. Another transcriptional regulator, Bmi-1, is required for the self-renewal capacity of HSCs, and in its absence animals die within 2 months of birth because of the failure to repopulate their red and white blood TABLE 2-1 Concentration and frequency of cells in human blood Cell type Cells/mm3 Red blood cells 5.0  106 Platelets 2.5  105 Leukocytes 7.3  103 Total leukocytes (%) Neutrophil 3.7–5.1  103 50–70 Lymphocyte 1.5–3.0  103 20–40 Monocyte 1–4.4  102 1–6 Eosinophil 2 1–2.2  10 1–3 Basophil 1.3  102 1 cells. Ikaros and Notch are both families of transcriptional regulators that have more specific effects on hematopoiesis. Ikaros is required for lymphoid but not myeloid development; animals survive in its absence but cannot mount a full immune response (i.e., they are severely immunocompromised). Notch1, one of four Notch family members, regulates the choice between T and B lymphocyte lineages (see Chapter 9). More master regulators of lineage commitment during hematopoiesis continue to be identified. The rate of hematopoiesis, as well as the production and release of specific cell lineages, is also responsive to environmental changes experienced by an organism. For instance, infection can result in the release of cytokines that markedly enhance the development of myeloid cells, including neutrophils. Investigators have also recently shown that the release of mature cells from the bone marrow is responsive to circadian cycles and regulated by the sympathetic nervous system. Distinguishing Blood Cells Early investigators originally classified cells based on their appearance under a microscope, often with the help of dyes. Their observations were especially helpful in distinguishing myeloid from lymphoid lineages, granulocytes from macrophages, neutrophils from basophils and eosinophils. Hematoxylin and eosin (H&E) stains are still commonly used in combination to distinguish cell types in blood smears and tissues. They highlight intracellular differences because of their pH sensitivity and different affinities for charged macromolecules in a cell. Thus, the basic dye hematoxylin binds basophilic nucleic acids, staining them blue, and the acidic dye eosin binds eosinophilic proteins in granules and cytoplasm, staining them pink. Microscopists drew astute inferences about cell function by detailed examination of the structure stained cells, as well as the behavior of live cells in solution. The advent of the flow cytometer in the 1980s revolutionized our understanding of cell subtypes by allowing us to evaluate multiple surface and internal proteins expressed by individual cells simultaneously. The development of ever more sophisticated fluorescent microscopy approaches to observe live cells in vitro and in vivo have allowed investigators to penetrate the complexities of the immune response in time and space. These advances coupled with abilities to manipulate cell function genetically have also revealed a remarkable diversity of cell types among myeloid and lymphoid cells, and continue to expose new functions and unexpected relationships among hematopoietic cells. Therefore, while our understanding of the cell subtypes is impressive, it is by no means complete. Cells of the Myeloid Lineage Are the First Responders to Infection Cells that arise from a common myeloid progenitor (CMP) include red blood cells (erythroid cells) as well as various types of white blood cells (myeloid cells such as granulocytes, monocytes, macrophages, and some dendritic cells). Myeloid Cells, Organs, and Microenvironments of the Immune System cells are the first to respond to the invasion of a pathogen and communicate the presence of an insult to cells of the lymphoid lineage (below). As we will see in Chapter 15, they also contribute to inflammatory diseases (asthma and allergy). Granulocytes Granulocytes are at the front lines of attack during an immune response and are considered part of the innate immune system. Granulocytes are white blood cells (leukocytes) that are classified as neutrophils, basophils, mast cells, or eosinophils on the basis of differences in cellular morphology and the staining of their characteristic cytoplasmic granules (Figure 2-2). All granulocytes have multilobed nuclei that make them visually distinctive and easily distinguishable from lymphocytes, whose nuclei are round. The cytoplasm of all granulocytes is replete with granules that are released in response to contact with pathogens. These granules contain a variety of proteins with distinct functions: Some damage pathogens directly; some regulate trafficking and activity of other white blood cells, including lymphocytes; and some contribute to the remodeling of tissues at the site of infection. See Table 2-2 for a partial list of granule proteins and their functions. Neutrophils constitute the majority (50% to 70%) of circulating leukocytes (see Figure 2-2a) and are much more numerous than eosinophils (1%–3%), basophils (1%), or mast cells (1%). After differentiation in the bone marrow, neutrophils are released into the peripheral blood and circulate for 7 to 10 hours before migrating into the tissues, where they have a life span of only a few days. In response to many types of infections, the number of circulating neutrophils increases significantly and more are recruited to tissues, partially in response to cues the bone marrow receives to produce and release more myeloid cells. The resulting transient increase in the number of circulating neutrophils, called leukocytosis, is used medically as an indication of infection. Neutrophils are recruited to the site of infection in response to inflammatory molecules (e.g., chemokines) generated by innate cells (including other neutrophils) that have engaged a pathogen. Once in tissues, neutrophils phagocytose (engulf) bacteria very effectively, and also secrete a range of proteins that have antimicrobial effects and tissue remodeling potential. Neutrophils are the dominant first responders to infection and the main cellular components of pus, where they accumulate at the end of their short lives. Although once considered a simple and “disposable” effector cell, the neutrophil has recently inspired renewed interest from investigations indicating that it may also regulate the adaptive immune response. Basophils are nonphagocytic granulocytes (see Figure 2-2b) that contain large granules filled with basophilic proteins (i.e., they stain blue in standard H&E staining protocols). Basophils are relatively rare in the circulation, but can be very potent. In response to binding of circulating antibodies, basophils release the contents of their granules. Histamine, one of the best known proteins in basophilic granules, increases blood vessel permeability and smooth muscle activity. Basophils (and eosinophils, below) are critical to | CHAPTER 2 33 our response to parasites, particularly helminths (worms), but in areas where worm infection is less prevalent, histamines are best appreciated as the cause of allergy symptoms. Like neutrophils, basophils may also secrete cytokines that modulate the adaptive immune response. Mast cells (see Figure 2-2c) are released from the bone marrow into the blood as undifferentiated cells; they mature only after they leave the blood. Mast cells can be found in a wide variety of tissues, including the skin, connective tissues of various organs, and mucosal epithelial tissue of the respiratory, genitourinary, and digestive tracts. Like circulating basophils, these cells have large numbers of cytoplasmic granules that contain histamine and other pharmacologically active substances. Mast cells also play an important role in the development of allergies. Basophils and mast cells share many features and their relationship is not unequivocally understood. Some speculate that basophils are the blood-borne version of mast cells; others speculate that they have distinct origins and functions. Eosinophils, like neutrophils, are motile phagocytic cells (see Figure 2-2d) that can migrate from the blood into the tissue spaces. Their phagocytic role is significantly less important than that of neutrophils, and it is thought that they play their most important role in the defense against multicellular parasitic organisms, including worms. They can be found clustering around invading worms, whose membranes are damaged by the activity of proteins released from eosinophilic granules. Like neutrophils and basophils, eosinophils may also secrete cytokines that regulate B and T lymphocytes, thereby influencing the adaptive immune response. In areas where parasites are less of a health problem, eosinophils are better appreciated as contributors to asthma and allergy symptoms. Myeloid Antigen-Presenting Cells Myeloid progenitors also give rise to a group of phagocytic cells (monocytes, macrophages, and dendritic cells) that have professional antigen-presenting cell (APC) function (Figure 2-3). Myeloid APCs are considered cellular bridges between the innate and adaptive immune systems because they make contact with a pathogen at the site of infection and communicate this encounter to T lymphocytes in the lymph node (“antigen presentation”). Each APC can respond to pathogens and secrete proteins that attract and activate other immune cells. Each can ingest pathogens via phagocytosis, digest pathogenic proteins into peptides, then present these peptide antigens on their membrane surfaces. Each can be induced to express a set of costimulatory molecules required for optimal activation of T lymphocytes. However, it is likely that each plays a distinct role during the immune response, depending on its locale and its ability to respond to pathogens. Dendritic cells, in particular, play a primary role in presenting antigen to—and activating—naïve T cells. Macrophages and neutrophils are especially efficient in removing both pathogen and damaged host cells, and can provide a first line of defense against pathogens. 34 PA R T I | Introduction (a) Neutrophil Multilobed nucleus Granules Phagosome (b) Basophil Glycogen Granule (c) Mast cell Granule (d) Eosinophil Granule FIGURE 2-2 Examples of granulocytes. (a, b, c, d) Hematoxylin and eosin (H&E) stains of indicated cells in blood smears. (a, middle) Neutrophil engulfing bacteria visualized by scanning electron microscopy (SEM) and colorized digitally. (b, middle) SEM of activated granulocytes (colorized). Each image is accompanied by a cartoon depicting the typical morphology of the indicated granulocyte. Note differences in the shape of the nucleus and in the number, color, and shape of the cytoplasmic granules. [2-2a, left: Science Source/Getty Images; 2-2a, right: Creative Commons, http://es.wikipedia.org/wiki/Archivo:Neutrophil_with_ anthrax_copy.jpg; 2-2b, left: Dr. Gladden Willis/Visuals Unlimited, Inc.; 2-2b, right: Steve Gschmeissner/Photo Researchers; 2-2c, left: Courtesy Gwen V. Childs, Ph.D., University of Arkansas for Medical Sciences; 2-2d, left; Pathpedia.com.] Cells, Organs, and Microenvironments of the Immune System TABLE 2-2 | CHAPTER 2 35 Examples of proteins contained in neutrophil, eosinophil, and basophil granules Cell type Molecule in granule Examples Function Neutrophil Proteases Antimicrobial proteins Protease inhibitors Histamine Elastase, Collagenase Defensins, lysozyme 1-anti-trypsin Tissue remodeling Direct harm to pathogens Regulation of proteases Vasodilation, inflammation Eosinophil Cationic proteins Ribonucleases Cytokines Chemokines EPO MBP ECP, EDN IL-4, IL-10, IL-13, TNF RANTES, MIP-1 Induces formation of ROS Vasodilation, basophil degranulation Antiviral activity Modulation of adaptive immune responses Attract leukocytes Basophil/Mast Cell Cytokines Lipid mediators Histamine IL-4, IL-13 Leukotrienes Modulation of adaptive immune response Regulation of inflammation Vasodilation, smooth muscle activation Monocytes make up about 5% to 10% of white blood cells and are a heterogeneous group of cells that migrate into tissues and differentiate into a diverse array of tissue-resident phagocytic cells, including macrophages and dendritic cells (see Figure 2-3a). During hematopoiesis in the bone marrow, granulocyte-monocyte progenitor cells differentiate into promonocytes, which leave the bone marrow and enter the blood, where they further differentiate into mature monocytes. Two broad categories of monocytes have recently been identified. Inflammatory monocytes enter tissues quickly in response to infection. Patrolling monocytes, a smaller group of cells that crawl slowly along blood vessels, provide a reservoir for tissue-resident monocytes in the absence of infection, and may quell rather than initiate immune responses. Monocytes that migrate into tissues in response to infection can differentiate into specific tissue macrophages. Like monocytes, macrophages can play several different roles. Some macrophages are long-term residents in tissues and play an important role in regulating their repair and regeneration. Other macrophages participate in the innate immune response and undergo a number of key changes when they are stimulated by encounters with pathogens or tissue damage. These are referred to as inflammatory macrophages and play a dual role in the immune system as effective phagocytes that can contribute to the clearance of pathogens from a tissue, as well as antigen-presenting cells that can activate T lymphocytes. Osteoclasts in the bone, microglial cells in the central nervous system, and alveolar macrophages in the lung are tissue-specific examples of macrophages with these properties. Activated, inflammatory macrophages are more effective than resting ones in eliminating potential pathogens for several reasons: They exhibit greater phagocytic activity, an increased ability to kill ingested microbes, increased secretion of inflammatory and cytotoxic mediators, and the abil- ity to activate T cells. More will be said about the antimicrobial activities of macrophages in Chapter 5. Activated macrophages also function more effectively as antigenpresenting cells for helper T cells (TH cells), which, in turn regulate and enhance macrophage activity. Thus, macrophages and TH cells facilitate each other’s activation during the immune response. Many macrophages also express receptors for certain classes of antibody. If an antigen (e.g., a bacterium) is coated with the appropriate antibody, the complex of antigen and antibody binds to antibody receptors on the macrophage membrane more readily than antigen alone and phagocytosis is enhanced. In one study, for example, the rate of phagocytosis of an antigen was 4000-fold higher in the presence of specific antibody to the antigen than in its absence. Thus, an antibody is an example of an opsonin, a molecule that binds an antigen marking it for recognition by immune cells. The modification of particulate antigens with opsonins (which come in a variety of forms) is called opsonization, a term from the Greek that literally means “to supply food” or “make tasty.” Opsonization is traditionally described as a process that enhances phagocytosis of an antigen, but it serves multiple purposes that will be discussed in subsequent chapters. Although most of the antigen ingested by macrophages is degraded and eliminated, early experiments with radiolabeled antigens demonstrated the presence of antigen peptides on the macrophage membrane. Although the macrophage is a very capable antigen-presenting cell, the dendritic cell is considered the most efficient activator of naïve T cells. The discovery of the dendritic cell (DC) by Ralph Steinman in the mid 1970s resulted in awarding of the Nobel Prize in 2011. Dendritic cells are critical for the initiation of the immune response and acquired their name because they are covered with long membranous extensions that resemble the dendrites of nerve cells and extend and retract dynamically, (a) Monocyte Lysosome Nucleus Phagosome (b) Macrophage Phagosome Pseudopodia Phagosome Phagolysosome Lysosome (c) Dendritic cell Phagosome Processes Phagosome Lysosome (d) Megakaryocyte Platelets Megakaryocyte FIGURE 2-3 Examples of monocytes, macrophages, dendritic cells, and megakaryocytes. (a, d) H&E stain of blood smear. (b) H&E stain of tissue section. (b, middle) SEM of macrophage engulfing mycobacteria (colorized). (c) SEM micrograph. Each image is accompanied by a cartoon depicting the typical morphology of the indicated cell. Note that macrophages are five- to tenfold larger than monocytes and contain more organelles, especially lysosomes. [2-3a, left: Pathpedia.com; 2-3b, left: Courtesy Dr. Thomas Caceci, Virginia-Maryland Regional College of Veterinary Medicine; 2-3b, middle: SPL/Photo Researchers; 2-3c, left: David Scharf/ Photo Researchers; 2-3d, left: Carolina Biological Supply Co./Visuals Unlimited.] Cells, Organs, and Microenvironments of the Immune System increasing the surface area available for browsing lymphocytes. They are more diverse a population of cells than once was thought, and seem to arise from both the myeloid and lymphoid lineages of hematopoietic cells. The functional distinctions among these diverse cells are still being clarified and are likely critically important in tailoring immune responses to distinct pathogens and targeting responding cells to distinct tissues. Dendritic cells perform the distinct functions of antigen capture in one location and antigen presentation in another. Outside lymph nodes, immature forms of these cells monitor the body for signs of invasion by pathogens and capture intruding or foreign antigens. They process these antigens, then migrate to lymph nodes, where they present the antigen to naïve T cells, initiating the adaptive immune response. When acting as sentinels in the periphery, immature dendritic cells take on their cargo of antigen in three ways. They engulf it by phagocytosis, internalize it by receptor-mediated endocytosis, or imbibe it by pinocytosis. Indeed, immature dendritic cells pinocytose fluid volumes of 1000 to 1500 m3 per hour, a volume that rivals that of the cell itself. Through a process of maturation, they shift from an antigen-capturing phenotype to one that is specialized for presentation of antigen to T cells. In making the transition, some attributes are lost and others are gained. Lost is the capacity for phagocytosis and large-scale pinocytosis. However, the ability to present antigen increases significantly, as does the expression of costimulatory molecules that are essential for the activation of naïve T cells. After activation, dendritic cells abandon residency in peripheral tissues, enter the blood or lymphatic circulation, and migrate to regions of the lymphoid organs, where T cells reside, and present antigen. It is important to note that, although they share a name, follicular dendritic cells do not arise in bone marrow and have completely different functions from those described for the dendritic cells discussed above. Follicular dendritic cells do not function as antigen-presenting cells for TH-cell activation. These dendritic cells were named for their exclusive location in organized structures of the lymph node called lymph follicles, which are rich in B cells. As discussed in Chapters 12 and 14, the interaction of B cells with follicular dendritic cells is an important step in the maturation and diversification of B cells. It is clear that myeloid cells are not the only cells that can present antigen efficiently. As mentioned above, lymphoidderived dendritic cells are fully capable APCs. In addition, activated B lymphocytes can act as professional antigenpresenting cells. B cells can internalize antigen very efficiently via their antigen-specific receptor, and can process and present antigenic peptides at the cell surface. Activated B cells also express the full complement of costimulatory molecules that are required to activate T cells. By presenting antigen directly to T cells, B cells efficiently solicit help, in the form of cytokines, that induces their differentiation into memory cells, as well as into antibody-producing cells (plasma cells). | CHAPTER 2 37 Erythroid Cells Cells of the erythroid lineage—erythrocytes, or red blood cells—also arise from a common myeloid precursor (sometimes referred to as a common myeloid-erythroid precursor). They contain high concentrations of hemoglobin, and circulate through blood vessels and capillaries delivering oxygen to surrounding cells and tissues. Damaged red blood cells can also release signals (free radicals) that induce innate immune activity. In mammals, erythrocytes are anuclear; their nucleated precursors (erythroblasts) extrude their nuclei in the bone marrow. However, the erythrocytes of almost all nonmammalian vertebrates (birds, fish, amphibians, and reptiles) retain their nuclei. Erythrocyte size and shape vary considerably across the animal kingdom—the largest red blood cells can be found among some amphibians, and the smallest among some deer species. Megakaryocytes Megakaryocytes are large myeloid cells that reside in the bone marrow and give rise to thousands of platelets, very small cells (or cell fragments) that circulate in the blood and participate in the formation of blood clots. Although platelets have some of the properties of independent cells, they do not have their own nuclei. Cells of the Lymphoid Lineage Regulate the Adaptive Immune Response Lymphocytes (Figure 2-4) are the principal cell players in the adaptive immune response. They represent 20% to 40% of circulating white blood cells and 99% of cells in the lymph. Lymphocytes can be broadly subdivided into three major populations on the basis of functional and phenotypic differences: B lymphocytes (B cells), T lymphocytes (T cells), and natural killer (NK) cells. In humans, approximately a trillion (1012) lymphocytes circulate continuously through the blood and lymph and migrate into the tissue spaces and lymphoid organs. We briefly review the general characteristics and functions of each lymphocyte group and its subsets below. Lymphocytes are relatively nondescript cells that are very difficult to distinguish morphologically. T and B cells, in particular, appear identical under a microscope. We therefore rely heavily on the signature of surface proteins they express to differentiate among lymphocyte subpopulations. Surface proteins expressed by immune cells are often referred to by the cluster of differentiation (CD or cluster of designation) nomenclature. This nomenclature was established in 1982 by an international group of investigators who recognized that many of the new antibodies produced by laboratories all over the world (largely in response to the advent of monoclonal antibody technology) were seeing the same proteins. They therefore defined clusters of antibodies that appeared to be seeing the same protein and assigned a name— a cluster of differentiation or CD—to each group. Although originally designed to categorize the multiple antibodies, the 38 PA R T I | Introduction (b) Lymphocyte with red blood cells (a) Lymphocyte TH helper cell TC cytotoxic T cell B cell (c) Plasma cell Plasma cell (d) NK cell (e) CLP NK T B TH Natural killer (NK) cell FIGURE 2-4 Examples of lymphocytes. (a, c, d) H&E stain of blood smear showing typical lymphocyte. Note that naïve B cells and T cells look identical by microscopy. (b) Scanning electron micrograph of lymphocytes and red blood cells. Cartoons depicting the typical morphology of the cells indicated accompany each image (including three different lymphocytes that would all have the same appearance). Note the enlarged area of cytoplasm of the plasma cell, which is occupied by an extensive network of endoplasmic reticulum and Golgi—an indication of the cell’s dedication to CD nomenclature is now firmly associated with specific surface proteins found on cells of many types. Table 2-3 lists some common CD molecules found on human and mouse lymphocytes. Note that the shift from use of a “common” name to the more standard “CD” name can take place slowly. (For example, investigators often still refer to the pan-T cell marker as “Thy1” rather than CD90, and the costimulatory molecules as “B7-1” and “B7-2,” rather than CD80 and CD86.) Appendix 1 lists over three hundred CD markers expressed by immune cells. In addition to their CD surface signatures, each B or T cell also expresses an antigen-specific receptor (the B cell receptor (BCR) or the T cell receptor (TCR), respectively) on its surface. Although the populations of B cells and T cells express a remarkable diversity of antigen receptors (more than a billion), all receptors on an individual cell’s surface have identical structures and therefore have identical specificities for antigen. If a given lymphocyte divides to form two daughter cells, both daughters bear antigen receptors with antigen specificities identical to each other and to the parental cell from which they arose, and so will any descendants they produce. The resulting population of lymphocytes, all arising from the same founding lymphocyte, is a clone. TH1 TC T H2 TREG TH17 antibody secretion. The NK cell also has more cytoplasm than a naïve lymphocyte; this is full of granules that are used to kill target cells. (e) A branch diagram that depicts the basic relationship among the lymphocyte subsets described in the text. [2-4a, left: Fred Hossler/ Visuals Unlimited; 2-4b: Creative Commons, http://commons.wikimedia.org/ wiki/File:SEM_blood_cells.jpg; 2-4c, left: Benjamin Koziner/Phototake; 2-4d: Courtesy Ira Ames, Ph.D., Dept. Cell & Developmental Biology, SUNY-Upstate Medical University.] At any given moment, a human or a mouse will contain tens of thousands, perhaps a hundred thousand, distinct mature T- and B-cell clones, each distinguished by its own unique and identical cohort of antigen receptors. Mature B cells and T cells are ready to encounter antigen, but they are considered naïve until they do so. Contact with antigen induces naïve lymphocytes to proliferate and differentiate into both effector cells and memory cells. Effector cells carry out specific functions to combat the pathogen, while the memory cells persist in the host, and upon rechallenge with the same antigen mediate a response that is both quicker and greater in magnitude. The first encounter with antigen is termed a primary response, and the re-encounter a secondary response. B Lymphocytes The B lymphocyte (B cell) derived its letter designation from its site of maturation, in the bursa of Fabricius in birds; the name turned out to be apt, as bone marrow is its major site of maturation in humans, mice, and many other mammals. Mature B cells are definitively distinguished from other lymphocytes and all other cells by their synthesis and Cells, Organs, and Microenvironments of the Immune System TABLE 2-3 | CHAPTER 2 39 Common CD markers used to distinguish functional lymphocyte subpopulations CD designation Function B cell TH TC NK cell CD2 Adhesion molecule; signal transduction     CD3 Signal transduction element of T-cell receptor     CD4 Adhesion molecule that binds to class II MHC molecules; signal transduction   (usually)  (usually)  CD5 Unknown  (subset)    CD8 Adhesion molecule that binds to class I MHC molecules; signal transduction   (usually)  (usually) (variable) CD16 (Fc RIII) Low-affinity receptor for Fc region of IgG     CD19 Signal transduction; CD21 co-receptor     CD21 (CR2) Receptor for complement (C3d and Epstein-Barr virus)     CD28 Receptor for costimulatory B7 molecule on antigen-presenting cells     CD32 (Fc RII) Receptor for Fc region of IgG     CD35 (CR1) Receptor for complement (C3b)     CD40 Signal transduction     CD45 Signal transduction     CD56 Adhesion molecule     Synonyms are shown in parentheses. display of the B-cell receptor (BCR), a membrane-bound immunoglobulin (antibody) molecule that binds to antigen. Each B cell expresses a surface antibody with a unique specificity, and each of the approximately 1.5–3  105 molecules of surface antibody has identical binding sites for antigen. B lymphocytes also can improve their ability to bind antigen through a process known as somatic hypermutation and can generate antibodies of several different functional classes through a process known as class switching. Somatic hypermutation and class switching are covered in detail in Chapter 12. Ultimately, activated B cells differentiate into effector cells known as plasma cells (see Figure 2-4c). Plasma cells lose expression of surface immunoglobulin and become highly specialized for secretion of antibody. A single cell is capable of secreting from a few hundred to more than a thousand molecules of antibody per second. Plasma cells do not divide and, although some long-lived populations of plasma cells are found in bone marrow, many die within 1 or 2 weeks. T Lymphocytes T lymphocytes (T cells) derive their letter designation from their site of maturation in the thymus. Like the B cell, the T cell expresses a unique antigen-binding receptor called the T-cell receptor. However, unlike membrane-bound antibodies on B cells, which can recognize soluble or particulate antigen, T-cell receptors only recognize processed pieces of antigen (typically peptides) bound to cell membrane proteins called major histocompatibility complex (MHC) molecules. MHC molecules are genetically diverse glycoproteins found on cell membranes (their structure and function are covered in detail in Chapter 8). The ability of MHC molecules to form complexes with antigen allows cells to decorate their surfaces with internal (foreign and self) proteins, exposing them to browsing T cells. MHC comes in two versions: class I MHC molecules, which are expressed by nearly all nucleated cells of vertebrate species, and class II MHC molecules, which are expressed by professional antigen-presenting cells and a few other cell types during inflammation. 40 PA R T I | Introduction T lymphocytes are divided into two major cell types—T helper (TH) cells and T cytotoxic (TC) cells—that can be distinguished from one another by the presence of either CD4 or CD8 membrane glycoproteins on their surfaces. T cells displaying CD4 generally function as TH cells and recognize antigen in complex with MHC class II, whereas those displaying CD8 generally function as TC cells and recognize antigen in complex with MHC class I. The ratio of CD4 to CD8 T cells is approximately 2:1 in normal mouse and human peripheral blood. A change in this ratio is often an indicator of immunodeficiency disease (e.g., HIV), autoimmune diseases, and other disorders. Naïve CD8 T cells browse the surfaces of antigenpresenting cells with their T-cell receptors. If and when they bind to an MHC-peptide complex, they become activated, proliferate, and differentiate into an effector cell called a cytotoxic T lymphocyte (CTL). The CTL has a vital function in monitoring the cells of the body and eliminating any cells that display foreign antigen complexed with class I MHC, such as virus-infected cells, tumor cells, and cells of a foreign tissue graft. To proliferate and differentiate optimally, naïve CD8 T cells also need help from mature CD4 T cells. Naïve CD4 T cells also browse the surfaces of antigenpresenting cells with their T-cell receptors. If and when they recognize an MHC-peptide complex, they can become activated and proliferate and differentiate into one of a variety of effector T cell subsets (see Figure 2-4e). T helper type 1 (TH1) cells regulate the immune response to intracellular pathogens, and T helper type 2 (TH2) cells regulate the response to many extracellular pathogens. Two additional TH cell subsets have been recently identified. T helper type 17 cells (TH17), so named because they secrete IL-17, play an important role in cell-mediated immunity and may help the defense against fungi. T follicular helper cells (TFH) play an important role in humoral immunity and regulate B-cell development in germinal centers. Which helper subtype dominates a response depends largely on what type of pathogen (intracellular versus extracellular, viral, bacterial, fungal, helminth) has infected an animal. Each of these CD4 T-cell subtypes produces a different set of cytokines that enable or “help” the activation of B cells, TC cells, macrophages, and various other cells that participate in the immune response. The network of cytokines that regulate and are produced by these effector cells is described in detail in Chapter 11. Another type of CD4 T cell, the regulatory T cell (TREG), has the unique capacity to inhibit an immune response. These cells can arise during maturation in the thymus from autoreactive cells (natural TREG), but also can be induced at the site of an immune response in an antigen-dependent manner (induced TREG). They are identified by the presence of CD4 and CD25 on their surfaces, as well as the expression of the internal transcription factor FoxP3. TREG cells are critical in helping us to quell autoreactive responses that have not been avoided via other mechanisms. In fact, mice depleted of TREG cells are afflicted with a constellation of destructive self-reactive inflammatory reactions. However, TREG cells may also play a role in limiting our normal T-cell response to a pathogen. CD4 and CD8 T-cell subpopulations may be even more diverse than currently described, and the field should expect identification of additional functional subtypes in the future. Natural Killer Cells Natural killer (NK) cells are lymphoid cells that are closely related to B and T cells. However, they do not express antigenspecific receptors and are considered part of the innate immune system. They are distinguished by the expression of a surface marker known as NK1.1, as well as the presence of cytotoxic granules. Once referred to as “large granular lymphocytes” because of their appearance under a microscope, NK cells constitute 5% to 10% of lymphocytes in human peripheral blood. They are efficient cell killers and attack a variety of abnormal cells, including some tumor cells and some cells infected with virus. They distinguish cells that should be killed from normal cells in a very clever way: by “recognizing” the absence of MHC class I, which is expressed by almost all normal cells, but is specifically down-regulated by some tumors and in response to some viral infections. How can cells recognize an absence? NK cells express a variety of receptors for self MHC class I that, when engaged, inhibit their ability to kill other cells. When NK cells encounter cells that have lost their MHC class I, these receptors are no longer engaged and can no longer inhibit the potent cytotoxic tendencies of the NK cell, which then releases its cytolytic granules and kills the abnormal target cell. NK cells also express receptors for immunoglobulins and can therefore decorate themselves with antibodies that bind pathogens or proteins from pathogens on the surface of infected cells. This allows an NK cell to make a connection with a variety of target cells (independently of their MHC class I expression). Once the antibodies bring the NK cell in contact with target cells, the NK cell releases its granules and induces cell death. The mechanism of NK-cell cytotoxicity, the focus of much current experimental study, is described further in Chapter 13. NKT Cells Another type of cell in the lymphoid lineage, NKT cells, have received a great deal of recent attention and share features with both conventional T lymphocytes and NK cells. Like T cells, NKT cells have T-cell receptors (TCRs), and some express CD4. Unlike most T cells, however, the TCRs of NKT cells are not very diverse and recognize specific lipids and glycolipids presented by a molecule related to MHC proteins known as CD1. Like their innate immune relatives, NK cells, NKT cells have antibody receptors, as well as other receptors classically associated with NK cells. Activated NKT cells can release cytotoxic granules that kill target cells, but they can also release large quantities of cytokines that can both enhance and suppress the immune response. They appear to be involved in human asthma, but also may inhibit the development of autoimmunity and cancer. Understanding the exact role of NKT cells in immunity is one research priority. Cells, Organs, and Microenvironments of the Immune System Primary Lymphoid Organs—Where Immune Cells Develop The ability of any stem cell to self-renew and differentiate depends on the structural organization and cellular function of specialized anatomic microenvironments known as stem cell niches. These sequestered regions are typically populated by a supportive network of stromal cells. Stem cell niche stromal cells express soluble and membrane-bound proteins that regulate cell survival, proliferation, differentiation, and trafficking. The organs that have microenvironments that support the differentiation of hematopoietic stem cells actually change over the course of embryonic development. However, by mid to late gestation, HSCs take up residence in the bone marrow, which remains the primary site of hematopoiesis throughout adult life. The bone marrow supports the maturation of all erythroid and myeloid cells and, in humans and mice, the maturation of B lymphocytes (as described in Chapter 10). HSCs are also found in blood and may naturally recirculate between the bone marrow and other tissues. This observation has simplified the process used to transplant blood cell progenitors from donors into patients who are deficient (e.g., patients who have undergone chemotherapy). Whereas once it was always necessary to aspirate bone marrow from the donor—a painful process that requires anesthesia—it is now sometimes possible to use enriched hematopoietic precursors from donor blood, which is much more easily obtained (see the Clinical Focus Box 2-2 on pages 42–43). Unlike B lymphocytes, T lymphocytes do not complete their maturation in the bone marrow. T lymphocyte precursors need to leave the bone marrow and travel to the unique microenvironments provided by the other primary lymphoid organ, the thymus, in order to develop into functional cells. The structure and function of the thymus will be discussed below and in more detail in Chapter 9. The Bone Marrow Provides Niches for Hematopoietic Stem Cells to Self-Renew and Differentiate into Myeloid Cells and B Lymphocytes The bone marrow is a primary lymphoid organ that supports self-renewal and differentiation of hematopoietic stem cells (HSCs) into mature blood cells. Although all bones contain marrow, the long bones (femur, humerus), hip bones (ileum), and sternum tend to be the most active sites of hematopoiesis. The bone marrow is not only responsible for the development and replenishment of blood cells, but it is also responsible for maintaining the pool of HSCs throughout the life of an adult vertebrate. The adult bone marrow (Figure 2-5), the paradigmatic adult stem cell niche, contains several cell types that coordinate HSC development, including (1) osteoblasts, versatile cells that both generate bone and control the differentiation | CHAPTER 2 41 of HSCs, (2) endothelial cells that line the blood vessels and also regulate HSC differentiation, (3) reticular cells that send processes connecting cells to bone and blood vessels, and, unexpectedly, (4) sympathetic neurons, which can control the release of hematopoietic cells from the bone marrow. A microscopic cross-section reveals that the bone marrow is tightly packed with stromal cells and hematopoietic cells at every stage of differentiation. With age, however, fat cells gradually replace 50% or more of the bone marrow compartment, and the efficiency of hematopoiesis decreases. The choices that an HSC makes depend largely on the environmental cues it receives. The bone marrow is packed with hematopoietic cells at all stages of development, but it is likely that the precursors of each myeloid and lymphoid subtype mature in distinct environmental micro-niches within the bone marrow. Our understanding of the microenvironments within the bone marrow that support specific stages of hematopoiesis is still developing. Evidence suggests, however, that the endosteal niche (the area directly surrounding the bone and in contact with bone-producing osteoblasts) and the vascular niche (the area directly surrounding the blood vessels and in contact with endothelial cells) play different roles (see Figure 2-5c). The endosteal niche appears to be occupied by quiescent HSCs in close association with osteoblasts that regulate stem cell proliferation. The vascular niche appears to be occupied by HSCs that have been mobilized to leave the endosteal niche to either differentiate or circulate. In addition, the more differentiated a cell is, the farther it appears to migrate from its supportive osteoblasts and the closer it moves to the more central regions of the bone. For example, the most immature B lymphocytes are found closest to the endosteum and osteoblasts, while the more mature B cells have moved into the more central sinuses of the bone marrow that are richly served by blood vessels. Finally, it is important to recognize that the bone marrow is not only a site for lymphoid and myeloid development but is also a site to which fully mature myeloid and lymphoid cells can return. Mature antibody-secreting B cells (plasma cells) may even take up long-term residence in the bone marrow. Whole bone marrow transplants, therefore, do not simply include stem cells but also include mature, functional cells that can both help and hurt the transplant effort. The Thymus Is a Primary Lymphoid Organ Where T Cells Mature T cell development is not complete until the cells undergo selection in the thymus (Figure 2-6). The importance of the thymus in T-cell development was not recognized until the early 1960s, when J.F.A.P. Miller, an Australian biologist, worked against the power of popular assumptions to advance his idea that the thymus was something other than a graveyard for cells. It was an underappreciated organ, very large in prepubescent animals, that was thought by some to be detrimental to an organism, and by others to be an evolutionary dead-end. The cells that populated it—small, thin-rimmed, 42 PA R T I | Introduction CLINICAL FOCUS Stem Cells—Clinical Uses and Potential Stem cell transplantation holds great promise for the regeneration of diseased, damaged, or defective tissue. Hematopoietic stem cells are already used to restore hematopoietic function, and their use in the clinic is described below. However, rapid advances in stem cell research have raised the possibility that other stem cell types may soon be routinely employed for replacement of a variety of cells and tissues. Two properties of stem cells underlie their utility and promise. They have the capacity to give rise to lineages of differentiated cells, and they are selfrenewing—each division of a stem cell creates at least one stem cell. If stem cells are classified according to their descent and developmental potential, three levels of stem cells can be recognized: pluripotent, multipotent, and unipotent. Pluripotent stem cells can give rise to an entire organism. A fertilized egg, the zygote, is an example of such a cell. In humans, the initial divisions of the zygote and its descendants produce cells that are also pluripotent. In fact, identical twins develop when pluripotent cells separate and develop into genetically identical fetuses. Multipotent stem cells arise from embryonic stem cells and can give rise to a more limited range of cell types. Further differentiation of multipotent stem cells leads to the formation of unipotent stem cells, which can generate only the same cell type as themselves. (Note that “pluripotent” is often used to describe the hematopoietic stem cell. Within the context of blood cell lineages this is arguably true; however, it is probably strictly accurate to call the HSC a multipotent stem cell.) Pluripotent cells, called embryonic stem cells, or simply ES cells, can be iso- lated from early embryos, and for many years it has been possible to grow mouse ES cells as cell lines in the laboratory. Strikingly, these cells can be induced to generate many different types of cells. Mouse ES cells have been shown to give rise to muscle cells, nerve cells, liver cells, pancreatic cells, and hematopoietic cells. Advances have made it possible to grow lines of human pluripotent stem cells and, most recently, to induce differentiated human cells to become pluripotent stem cells. These are developments of considerable importance to the understanding of human development, and they also have great therapeutic potential. In vitro studies of the factors that determine or influence the development of human pluripotent stem cells along specific developmental paths are providing considerable insight into how cells differentiate into specialized cell types. This research is driven in part by the great potential for using pluripotent stem cells to generate cells and tissues that could replace diseased or damaged tissue. Success in this endeavor would be a major advance because transplantation medicine now depends entirely on donated organs and tissues, yet the need far exceeds the number of donations, and the need is increasing. Success in deriving cells, tissues, and organs from pluripotent stem cells could provide skin replacement for burn patients, heart muscle cells for those with chronic heart disease, pancreatic islet cells for patients with diabetes, and neurons for the treatment of Parkinson’s disease or Alzheimer’s disease. The transplantation of HSCs is an important therapy for patients whose hematopoietic systems must be replaced. It has three major applications: featureless cells called thymocytes—looked dull and inactive. However, Miller proved that the thymus was the allimportant site for the maturation of T lymphocytes (see the Classic Experiment Box 2-3 on pages 46–47). T-cell precursors, which still retain the ability to give rise to multiple hematopoietic cell types, travel via the blood from the bone marrow to the thymus. Immature T cells, • Providing a functional immune system to individuals with a genetically determined immunodeficiency, such as severe combined immunodeficiency (SCID). • Replacing a defective hematopoietic system with a functional one to cure patients with life-threatening nonmalignant genetic disorders in hematopoiesis, such as sickle-cell anemia or thalassemia. • Restoring the hematopoietic system of cancer patients after treatment with doses of chemotherapeutic agents and radiation. This approach is particularly applicable to leukemias, including acute myeloid leukemia, which can be cured only by destroying the patient’s own hematopoietic system—the source of the leukemia cells. Clinicians also hope that this approach can be used to facilitate treatment of solid tumors. Highdose radiation and cytotoxic regimens can be much more effective at killing solid tumors than therapies using more conventional doses of cytotoxic agents; however, they destroy the immune system, and stem cell transplantation makes it possible to recover from such drastic treatment. Hematopoietic stem cells have extraordinary powers of regeneration. Experiments in mice indicate that as few as one HSC can completely restore the erythroid population and the immune system. In humans, for instance, as little as 10% of a donor’s total volume of bone marrow can provide enough HSCs to completely restore the recipient’s hematopoietic system. Once injected into a vein, HSCs enter the circulation and find their own way to known as thymocytes (thymus cells) because of their site of maturation, pass through defined developmental stages in specific thymic microenvironments as they mature into functional T cells. The thymus is a specialized environment where immature T cells generate unique antigen receptors (T cell receptors, or TCRs) and are then selected on the basis of their reactivity to self MHC-peptide complexes expressed Cells, Organs, and Microenvironments of the Immune System | CHAPTER 2 43 BOX 2-2 the bone marrow, where they begin the process of engraftment. In addition, HSCs can be preserved by freezing. This means that hematopoietic cells can be “banked.” After collection, the cells are treated with a cryopreservative, frozen, and then stored for later use. When needed, the frozen preparation is thawed and infused into the patient, where it reconstitutes the hematopoietic system. This cell-freezing technology even makes it possible for individuals to store their own hematopoietic cells for transplantation to themselves at a later time. Currently, this procedure is used to allow cancer patients to donate cells before undergoing chemotherapy and radiation treatments, then later reconstitute their hematopoietic system using their own cells. Transplantation of stem cell populations may be autologous (the recipient is also the donor), syngeneic (the donor is genetically identical; i.e., an identical twin of the recipient), or allogeneic (the donor and recipient are not genetically identical). In any transplantation procedure, genetic differences between donor and recipient can lead to immune-based rejection reactions. Aside from host rejection of transplanted tissue (host versus graft), lymphocytes conveyed to the recipient via the graft can attack the recipient’s tissues, thereby causing graftversus-host disease (GVHD), a life-threatening affliction. In order to suppress rejection reactions, powerful immunosuppressive drugs must be used. Unfortunately, these drugs have serious side effects, and immunosuppression increases the patient’s risk of infection and susceptibility to tumors. Consequently, HSC transplantation has the fewest complications when there is genetic identity between donor and recipient. At one time, bone marrow transplantation was the only way to restore the hematopoietic system. However, both peripheral blood and umbilical cord blood are now also common sources of hematopoietic stem cells. These alternative sources of HSCs are attractive because the donor does not have to undergo anesthesia or the highly invasive procedure used to extract bone marrow. Although peripheral blood may replace marrow as a major source of hematopoietic stem cells for many applications, bone marrow transplantation still has some advantages (e.g., marrow may include stem cell subsets that are not as prevalent in blood). To obtain HSCenriched preparations from peripheral blood, agents are used to induce increased numbers of circulating HSCs, and then the HSC-containing fraction is separated from the plasma and red blood cells in a process called leukapheresis. If necessary, further purification can be done to remove T cells and to enrich the CD34 population. Umbilical cord blood contains an unusually high frequency of hematopoietic stem cells. Furthermore, it is obtained from placental tissue (the “afterbirth”), which is normally discarded. Consequently, umbilical cord blood has become an attractive source of cells for HSC transplantation. For reasons that remain incompletely understood, cord blood stem cell transplants do not engraft as reliably as peripheral blood stem cell transplants; however, grafts of cord blood cells produce GVHD less frequently than marrow grafts, probably because cord blood has fewer mature T cells. Beyond its current applications in cancer treatment, autologous stem cell transplantation can also be useful for on the surface of thymic stromal cells. Those thymocytes whose T-cell receptors bind self MHC-peptide complexes with too high affinity are induced to die (negative selection), and those thymocytes that bind self MHC-peptides with an intermediate affinity undergo positive selection, resulting in their survival, maturation, and migration to the thymic medulla. Most thymocytes do not navigate the jour- gene therapy, the introduction of a normal gene to correct a disorder caused by a defective gene. One of the most highly publicized gene therapy efforts—the introduction of the adenosine deaminase (ADA) gene to correct a form of severe combined immunodeficiency (SCID)—was performed successfully on hematopoietic stem cells. The therapy entails removing a sample of hematopoietic stem cells from a patient, inserting a functional gene to compensate for the defective one, and then reinjecting the engineered stem cells into the donor. The advantage of using stem cells in gene therapy is that they are selfrenewing. Consequently, at least in theory, patients would have to receive only a single injection of engineered stem cells. In contrast, gene therapy with engineered mature lymphocytes or other blood cells would require periodic injections because these cells are not capable of self-renewal. In the case of the SCID patients, hematopoietic stem cells were successfully infected with a retrovirus engineered to express the ADA gene. The cells were returned to the patients and did, indeed, correct the deficiency. Patients who previously could not generate lymphocytes to protect themselves from infection were able to generate normal cells and live relatively normal lives. Unfortunately, in a number of patients, the retrovirus used to introduce the ADA gene integrated into parts of the genome that resulted in leukemia. Investigators continue to work to improve the safety and efficiency of gene delivery; more successful gene therapy efforts are clearly in the future. ney through the thymus successfully; in fact, it is estimated that 95% of thymocytes die in transit. The majority of cells die because they have too low an affinity for the self-antigenMHC combinations that they encounter on the surface of thymic epithelial cells and fail to undergo positive selection. These developmental events take place in several distinct thymic microenvironments (see Figure 2-6). T-cell precursors PA R T I 44 | Introduction (a) (b) Sternum Humerus Medullary cavity Endosteal niche Ileum Vascular niche Femur FIGURE 2-5 The bone marrow microenvironment. (a) Multiple bones support hematopoiesis, including the hip (ileum), femur, sternum, and humerus. (b) This figure shows a typical cross-section of a bone with a medullary (marrow) cavity. (c) Blood vessels (central sinus and medullary artery) run through the center of the bone and form a network of capillaries in close association with bone and bone surface (endosteum). Both the cells that line the blood vessels (endothelium) and the cells that line the bone (osteoblasts) generate niches that support hematopoietic stem cell (HSC) self-renewal and differentiation. The most immature cells appear to be associated with the endosteal (bone) niche; as they mature, they migrate toward the vascular (blood vessel) niche. Fully differentiated cells exit the marrow via blood vessels. [2-5b; Courtesy of Indiana University School of Medicine.] HSC (c) Reticular cells Hematopoietic stem cells (HSC) Osteoblasts Endosteal niche Vascular niche Artery Vein Medullary artery Medullary cavity Central sinus Sympathetic neuron | Cells, Organs, and Microenvironments of the Immune System CHAPTER 2 (b) (a) Thyroid Thymus Nerve Nerve Thymus Heart Diaphragm Bone marrow (c) Cortex Medulla FIGURE 2-6 The structure of the thymus. The thymus is found just above the heart (a, b) and is largest prior to puberty, when it begins to shrink. Panel (c) depicts a stained thymus tissue section and (d) a cartoon of the microenvironments: the cortex, which is densely populated with DP immature thymocytes (blue) and the medulla, which is sparsely populated with SP mature thymocytes. These major regions are separated by the corticomedullary junction (CMJ), where cells enter from and exit to the bloodstream. The area between the cortex and the thymic capsule, the subcapsular cortex, is a site of much proliferation of the youngest (DN) thymocytes. The route taken by a typical thymocyte during its development from the DN to DP to SP stages is shown. Thymocytes are positively selected in the cortex. Autoreactive thymocytes are negatively selected in the medulla; some may also be negatively selected in the cortex. [2-6c: Dr. Gladden Willis/Getty Images.] (d) Thymus Capsule DN Subcapsular cortex DP Cortex SP DN SP Cortical medullary junction (CMJ) SP Medulla 45 46 PA R T I | Introduction CLASSIC EXPERIMENT The Discovery of a Thymus—and Two J.F.A.P. MILLER DISCOVERED THE FUNCTION OF THE THYMUS In 1961, Miller, who had been investigating the thymus’s role in leukemia, published a set of observations in The Lancet that challenged notions that, at best, this organ served as a cemetery for lymphocytes and, at worst, was detrimental to health (Figure 1). He noted that when this organ was removed in very young mice (in a process known as thymectomy), the subjects became susceptible to a variety of infections, failed to reject skin grafts, and died prematurely. On close examination of their circulating blood cells, they also appeared to be missing a type of cell that another investigator, James Gowans, had associated with cellular and humoral immune responses. Miller concluded that the thymus produced functional immune cells. Several influential investigators could not repeat the data and questioned Miller’s conclusions. Some speculated that the mouse strain he used was peculiar, others that his mice were exposed to too many pathogens and their troubles were secondary to infection. Dr. Miller responded to each of these criticisms experimentally, assessing the impact of thymectomy in different mouse strains and in germ-free facilities. His results were unequivocal, and his contention that this organ generated functional lymphocytes was vindicated. Elegant experiments by Dr. Miller, James Gowans, and others subsequently showed that the thymus produced a different type of lymphocyte than the bone marrow. This cell did not produce antibodies directly, but, instead, was required for optimal antibody production. It was called a T cell after the thymus, its organ of origin. Immature T cells are known as thymocytes. Miller is one of the few scientists credited with the discovery of the function of an entire organ. A SECOND THYMUS No one expected a new anatomical discovery in immunology in the 21st century. However, in 2006 Hans-Reimer Rodewald and his colleagues reported the existence of a second thymus in mice. The conventional thymus is a bi-lobed organ that sits in the thorax right above the heart. Rodewald and his colleagues discovered thymic tissue that sits in the neck, near the cervical vertebrae, of mice. This cervical FIGURE 1 J.F.A.P. Miller (above) in 1961 and the first page (opposite) of his Lancet article (1961) describing his discovery of the function of the thymus. [The Walter and Eliza Hall Institute of Medical Research.] thymic tissue is smaller in mass than the conventional thymus, consists of a single lobe or clusters of single lobes, and is populated by relatively more mature thymocytes. However, it contributes to T-cell development very effectively and clearly contributes to the mature T-cell repertoire. Rodewald’s findings raise the possibility that some of our older observations and assumptions about thymic function need to be reexamined. In particular, studies based on thymectomy that indicated T cells could develop outside the thymus may need to be reassessed. The cells found may have come from this more obscure but functional thymic tissue. The evolutionary implications of this thymus are also interesting—thymi are found in the neck in several species, including the koala and kangaroo. REFERENCES Miller, J. F. 1961. Immunological function of the thymus. Lancet 2:748–749. Miller, J.F.A.P. 2002. The discovery of thymus function and of thymus-derived lymphocytes. Immunological Reviews 185:7–14. Rodewald, H-R. 2006. A second chance for the thymus. Nature 441:942–943. Cells, Organs, and Microenvironments of the Immune System | CHAPTER 2 47 BOX 2-3 [Immunological Function of the Thymus, J.F.A.P. Miller, The Lancet, Elsevier 30 September 1961 © 1961, Elsevier] 48 PA R T I | Introduction enter the thymus in blood vessels at the corticomedullary junction between the thymic cortex, the outer portion of the organ, and the thymic medulla, the inner portion of the organ. At this stage thymocytes express neither CD4 nor CD8, markers associated with mature T cells. They are therefore called double negative (DN) cells. DN cells first travel to the region under the thymic capsule, a region referred to as the subcapsular cortex, where they proliferate and begin to generate their T-cell receptors. Thymocytes that successfully express TCRs begin to express both CD4 and CD8, becoming double positive (DP) cells, and populate the cortex, the site where most (85% or more) immature T cells are found. The cortex features a distinct set of stromal cells, cortical thymic epithelial cells (cTECs), whose long processes are perused by thymocytes testing the ability of their T-cell receptors to bind MHC-peptide complexes (Video 2-1). Thymocytes that survive selection move to the thymic medulla, where positively selected thymocytes encounter specialized stromal cells, medullary thymic epithelial cells (mTECs). Not only do mTECs support the final steps of thymocyte maturation, but they also have a unique ability to express proteins that are otherwise found exclusively in other organs. This allows them to negatively select a group of potentially very damaging, autoreactive T cells that could not be deleted in the cortex.1 Mature thymocytes, which express only CD4 or CD8 and are referred to as single positive (SP), leave the thymus as they entered: via the blood vessels of the corticomedullary junction. Maturation is finalized in the periphery, where these new T cells (recent thymic emigrants) explore antigens presented in secondary lymphoid tissue, including spleen and lymph nodes. Secondary Lymphoid Organs—Where the Immune Response Is Initiated As just described, lymphocytes and myeloid cells develop to maturity in the primary lymphoid system: T lymphocytes in the thymus, and B cells, monocytes, dendritic cells, and granulocytes in the bone marrow. However, they encounter antigen and initiate an immune response in the microenvironments of secondary lymphoid organs (SLOs). Secondary Lymphoid Organs Are Distributed Throughout the Body and Share Some Anatomical Features Lymph nodes and the spleen are the most highly organized of the secondary lymphoid organs and are compartmentalized from the rest of the body by a fibrous capsule. A some- 1 Note that some investigators describe positive selection as taking place in the cortex and negative selection solely in the medulla. However, several lines of evidence suggest that negative selection can also occur in the cortex, and we have adopted this perspective for this text. what less organized system of secondary lymphoid tissue, collectively referred to as mucosa-associated lymphoid tissue (MALT), is found associated with the linings of multiple organ systems, including the gastrointestinal (GI) and respiratory tracts. MALT includes tonsils, Peyer’s patches (in the small intestine), and the appendix, as well as numerous lymphoid follicles within the lamina propria of the intestines and in the mucous membranes lining the upper airways, bronchi, and genitourinary tract (Figure 2-7). Although secondary lymphoid organs vary in their location and degree of organization, they share key features. All SLOs include anatomically distinct regions of T-cell and B-cell activity, and all develop lymphoid follicles, which are highly organized microenvironments that are responsible for the development and selection of B cells that produce high-affinity antibodies. Lymphoid Organs Are Connected to Each Other and to Infected Tissue by Two Different Circulatory Systems: Blood and Lymphatics The immune cells are the most mobile cells in a body and use two different systems to traffic through tissues: the blood system and the lymphatic system. The blood has access to virtually every organ and tissue and is lined by endothelial cells that are very responsive to inflammatory signals. Hematopoietic cells can transit through the blood system—away from the heart via active pumping networks (arteries) and back to the heart via passive valve-based systems (veins) within minutes. Most lymphocytes enter secondary lymphoid organs via specialized blood vessels, and leave via the lymphatic system. The lymphatic system is a network of thin walled vessels that play a major role in immune cell trafficking, including the travel of antigen and antigen-presenting cells to secondary lymphoid organs and the exit of lymphocytes from lymph nodes. Lymph vessels are filled with a protein-rich fluid (lymph) derived from the fluid component of blood (plasma) that seeps through the thin walls of capillaries into the surrounding tissue. In an adult, depending on size and activity, seepage can add up to 2.9 liters or more during a 24-hour period. This fluid, called interstitial fluid, permeates all tissues and bathes all cells. If this fluid were not returned to the circulation, the tissue would swell, causing edema that would eventually become life threatening. We are not afflicted with such catastrophic edema because much of the fluid is returned to the blood through the walls of venules. The remainder of the interstitial fluid enters the delicate network of primary lymphatic vessels. The walls of the primary vessels consist of a single layer of loosely apposed endothelial cells. The porous architecture of the primary vessels allows fluids and even cells to enter the lymphatic network. Within these vessels, the fluid, now called lymph, flows into a series of progressively larger collecting vessels called lymphatic vessels (see Figures 2-7b and 2-7c). Cells, Organs, and Microenvironments of the Immune System | CHAPTER 2 49 (b) (a) Adenoids Right lymphatic duct Right subclavian vein Thymus Thoracic duct Left subclavian vein Lymph nodes Spleen Peyer's patches Bone marrow Lymph node Lymphatic vessel (c) Smooth muscle Lymphatic capillary Arteriole Tissue lymphatics Blood capillaries Interstitial fluid Venule Lymph FIGURE 2-7 The human lymphoid system. The primary organs (bone marrow and thymus) are shown in red; secondary organs and tissues, in blue. These structurally and functionally diverse lymphoid organs and tissues are interconnected by the blood vessels (not shown) and lymphatic vessels (purple). Most of the body’s lymphatics eventually drain into the thoracic duct, which empties into the left subclavian vein. However, the vessels draining the right arm and right side of the head (shaded blue) converge to form the right lymphatic duct, which empties into the right subclavian vein. The inset (b) shows the lymphatic vessels in more detail, and (c) shows the relationship between blood and lymphatic capillaries in tissue. The lymphatic capillaries pick up interstitial fluid, particulate and soluble proteins, as well as immune cells from the tissue surrounding the blood capillaries (see arrows). [Part (a): Adapted from H. Lodish et al., 1995, Molecular Cell Biology, 3rd ed., Scientific American Books, New York.] All cells and fluid circulating in the lymph are ultimately returned to the blood system. The largest lymphatic vessel, the thoracic duct, empties into the left subclavian vein. It collects lymph from all of the body except the right arm and right side of the head. Lymph from these areas is collected into the right lymphatic duct, which drains into the right subclavian vein (see Figure 2-7a). By returning fluid lost from the blood, the lymphatic system ensures steady-state levels of fluid within the circulatory system. The heart does not pump the lymph through the lymphatic system; instead, the slow, low-pressure flow of lymph is achieved by the movements of the surrounding muscles. Therefore, activity enhances lymph circulation. Importantly, a series of one-way valves along the lymphatic vessels ensures that lymph flows in only one direction. When a foreign antigen gains entrance to the tissues, it is picked up by the lymphatic system (which drains all the tissues of the body) and is carried to various organized 50 PA R T I | Introduction lymphoid tissues such as lymph nodes, which trap the foreign antigen. Antigen-presenting cells that engulf and process the antigen also can gain access to lymph. In fact, as lymph passes from the tissues to lymphatic vessels, it becomes progressively enriched in specific leukocytes, including lymphocytes, dendritic cells, and macrophages. Thus, the lymphatic system also serves as a means of transporting white blood cells and antigen from the connective tissues to organized lymphoid tissues, where the lymphocytes can interact with the trapped antigen and undergo activation. Most secondary lymphoid tissues are situated along the vessels of the lymphatic system. The spleen is an exception and is served only by blood vessels. All immune cells that traffic through tissues, blood, and lymph nodes are guided by small molecules known as chemokines. These proteins are secreted by stromal cells, antigenpresenting cells, lymphocytes, and granulocytes, and form gradients that act as attractants and guides for other immune cells, which express an equally diverse set of receptors for these chemokines. The interaction between specific chemokines and cells expressing specific chemokine receptors allows for a highly refined organization of immune cell movements. The Lymph Node Is a Highly Specialized Secondary Lymphoid Organ Lymph nodes (Figure 2-8) are the most specialized SLOs. Unlike the spleen, which also regulates red blood cell flow and fate, lymph nodes are fully committed to regulating an immune response. They are encapsulated, bean-shaped structures that include networks of stromal cells packed with lymphocytes, macrophages, and dendritic cells. Connected to both blood vessels and lymphatic vessels, lymph nodes are the first organized lymphoid structure to encounter antigens that enter the tissue spaces. The lymph node provides ideal microenvironments for encounters between antigen and lymphocytes and productive, organized cellular and humoral immune responses. Structurally, a lymph node can be divided into three roughly concentric regions: the cortex, the paracortex, and the medulla, each of which supports a distinct microenvironment (see Figure 2-8). The outermost layer, the cortex, contains lymphocytes (mostly B cells), macrophages, and follicular dendritic cells arranged in follicles. Beneath the cortex is the paracortex, which is populated largely by T lymphocytes and also contains dendritic cells that migrated from tissues to the node. The medulla is the innermost layer, and the site where lymphocytes exit (egress) the lymph node through the outgoing (efferent) lymphatics. It is more sparsely populated with lymphoid lineage cells, which include plasma cells that are actively secreting antibody molecules. Antigen travels from infected tissue to the cortex of the lymph node via the incoming (afferent) lymphatic vessels, which pierce the capsule of a lymph node at numerous sites and empty lymph into the subcapsular sinus (see Figure 2-8b). It enters either in particulate form or is processed and presented as peptides on the surface of migrating antigenpresenting cells. Particulate antigen can be trapped by resident antigen-presenting cells in the subcapsular sinus or cortex, and it can be passed to other antigen-presenting cells, including B lymphocytes. Alternatively, particulate antigen can be processed and presented as peptide-MHC complexes on cell surfaces of resident dendritic cells that are already in the T-cell-rich paracortex. T Cells in the Lymph Node It takes every naïve T lymphocyte about 16 to 24 hours to browse all the MHC-peptide combinations presented by the antigen-presenting cells in a single lymph node. Naïve lymphocytes enter the cortex of the lymph node by passing between the specialized endothelial cells of high endothelial venules (HEV), so-called because they are lined with unusually tall endothelial cells that give them a thickened appearance (Figure 2-9a). Once naïve T cells enter the lymph node, they browse MHC-peptide antigen complexes on the surfaces of the dendritic cells present in the paracortex. The paracortex is traversed by a web of processes that arise from stromal cells called fibroblast reticular cells (FRCs) (Figure 2-9b). This network is referred to as the fibroblast reticular cell conduit system (FRCC) and guides T-cell movements via associated adhesion molecules and chemokines. Antigen-presenting cells also appear to wrap themselves around the conduits, giving circulating T cells ample opportunity to browse their surfaces as they are guided down the network. The presence of this specialized network elegantly enhances the probability that T cells will meet their specific MHC-peptide combination. T cells that browse the lymph node but do not bind MHC-peptide combinations exit not via the blood, but via the efferent lymphatics in the medulla of the lymph node (see Figure 2-8). T cells whose TCRs do bind to an MHC-peptide complex on an antigen-presenting cell that they encounter in the lymph node will stop migrating and take up residence in the node for several days. Here it will proliferate and, depending on cues from the antigen-presenting cell itself, its progeny will differentiate into effector cells with a variety of functions. CD8 T cells gain the ability to kill target cells. CD4 T cells can differentiate into several different kinds of effector cells, including those that can further activate macrophages, CD8 T cells, and B cells. B Cells in the Lymph Node The lymph node is also the site where B cells are activated and differentiate into high-affinity antibody-secreting plasma cells. B cell activation requires both antigen engagement by the B-cell receptor (BCR) and direct contact with an activated CD4 TH cell. Both events are facilitated by the anatomy of the lymph node. Like T cells, B cells circulate through the blood and lymph and visit the lymph nodes on a daily basis, entering via the HEV. They respond to specific signals and chemokines that draw them not to the paracortex but to the lymph node follicle. Although they may initially take advantage of the FRCC for guidance, they Cells, Organs, and Microenvironments of the Immune System (a) CHAPTER 2 51 (b) Follicle (B cell zone) Adenoids Left subclavian vein Right lymphatic duct | Paracortex (T cell zone) Incoming (afferent) lymphatics Lymph nodes Antigen, APC Right subclavian vein Medulla Thymus Spleen Outer cortex Peyer's patches Blood vessels Bone marrow Thoracic duct Outgoing (efferent) lymphatics Lymphocytes, APC Naïve lymphocytes (d) (c) B-cell follicle T-cell zone: paracortex Follicle Germinal center Tissue lymphatics FIGURE 2-8 Structure of a lymph node. The microenvironments of the lymph node support distinct cell activities. (a) The lymph nodes are dispersed throughout the body and are connected by lymphatic vessels as well as blood vessels (not shown). (b) A drawing of the major features of a lymph node shows the major vessels that serve the organ: incoming (afferent) and outgoing (efferent) lymphatic vessels, and the arteries and veins. It also depicts the three major tissue layers: the outer cortex, the paracortex, and the innermost region, the medulla. Macrophages and dendritic cells, which trap antigen, are present in the cortex and paracortex. T cells are concentrated in the paracortex; B cells are primarily in the cortex, within follicles and germinal centers. The medulla is populated largely by antibody-producing plasma cells and is the site where cells exit via the efferent lymphatics. Naïve lymphocytes circulating in the blood enter the node via high endothelial venules (HEV) via a process called extravasation (see Advances Box 14-2). Antigen and some leukocytes, including antigenpresenting cells, enter via afferent lymphatic vessels. All cells exit via efferent lymphatic vessels. (c) This stained tissue section shows the cortex with a number of ovoid follicles, which is surrounded by the T-cell-rich paracortex. (d) A stained lymph node section showing a follicle that includes a germinal center (otherwise referred to as a secondary follicle). [2-8c: Dr. Gladden Willis/Getty Images; 2-8d: Image Source/Alamy.] ultimately depend upon follicular dendritic cells (FDCs) for guidance (Figure 2-9c). FDCs are centrally important in maintaining follicular and germinal center structure and “presenting” antigen to differentiating B cells. B cells differ from T cells in that their receptors can recognize free antigen. A B cell will typically meet its antigen in the follicle. If its BCR binds to antigen, the B cell becomes partially activated and engulfs and processes that antigen. As mentioned above, B cells, in fact, are specialized antigen-presenting cells that present processed peptideMHC complexes on their surface to CD4 TH cells. Recent data show that B cells that have successfully engaged and processed antigen change their migration patterns and move to the T-cell-rich paracortex, where they increase their chances of encountering an activated CD4 TH cell that will recognize the MHC-antigen complex they present. When they successfully engage this TH cell, they maintain contact for a number of hours, becoming fully activated and receiving signals that induce B cell proliferation (see Chapter 14). Some activated B cells differentiate directly into an antibody-producing cell (plasma cell) but others re-enter the follicle to establish a germinal center. A follicle that develops a germinal center is sometimes referred to as a secondary follicle; a follicle without a germinal center is sometimes referred to as a primary follicle. 52 PA R T I | Introduction (a) Afferent lymphatic vessel high endothelial venule Lymphatic endothelial cell Blood endothelial cell DC Perivascular sheath Basal lamina Soluble molecule Particulate antigen Naïve B cell Naïve T cell (b) Follicular reticular cell conduit system FRC T cell DC HEV (c) Follicular dendritic cell FDC B cell FIGURE 2-9 Features of lymph node microenvironments. The lymph node microenvironments are maintained and regulated by distinct cell types and structures. (a) The afferent lymphatics are the vessels through which dendritic cells, and particulate and soluble antigen, enter the lymph node. The high endothelial venules (HEVs) are the vessels through which naïve T and B cells enter the lymph node (via extravasation). (b) The paracortex is crisscrossed by processes and conduits formed by fibroblastic reticular cells (FRCs), which guide the migration of antigen- nie Favre and Sanjiv A. Luther, University of Lausanne, Switzerland. 2-9c, left: Photograph courtesy of Mohey Eldin M. El Shikh.] Germinal centers are remarkable substructures that facilitate the generation of B cells with increased receptor affinities. In the germinal center, an antigen-specific B cell clone will proliferate and undergo somatic hypermutation of the genes coding for their antigen receptors. Those receptors that retain the ability to bind antigen with the highest affinity survive and differentiate into plasma cells that travel to the medulla of the lymph node. Some will stay and release antibodies into the bloodstream; others will exit through the efferent lymphatics and take up residence in the bone marrow, where they will continue to release antibodies into circulation. The initial activation of B cells and establishment of the germinal center take place within 4 to 7 days of the initial infection, but germinal centers remain active for 3 weeks or more (Chapter 12). Lymph nodes swell visibly and sometimes painfully, particularly during those first few days after infection. This swelling is due both to an increase in the number of lymphocytes induced to migrate into the node as well as the proliferation of antigen-specific T and B lymphocytes within the lobe. presenting cells and T cells, facilitating their interactions. The left panel shows an immunofluorescence microscopy image with the FRC shown in red and T cells in green. The right panel shows a cartoon of the network and cell participants. (c) The B-cell follicle contains a network of follicular dendritic cells (FDCs), which are shown as an SEM image as well as a cartoon. FDCs guide the movements and interactions of B cells. [2-9b, left: Courtesy of Stepha- Cells, Organs, and Microenvironments of the Immune System The Generation of Memory T and B Cells in the Lymph Node The interactions between TH cells and APCs, and between activated TH cells and activated B cells, results not only in the proliferation of antigen-specific lymphocytes and their functional differentiation, but also in the generation of memory T and B cells. Memory T and B cells can take up residence in secondary lymphoid tissues or can exit the lymph node and travel to and among tissues that first encountered the pathogen. Memory T cells that reside in secondary lymphoid organs are referred to as central memory cells and are distinct in phenotype and functional potential from effector memory T cells that circulate among tissues. Memory cell phenotype, locale, and activation requirements are an active area of investigation and will be discussed in in more detail in Chapters 11 and 12. The Spleen Organizes the Immune Response Against Blood-Borne Pathogens The spleen, situated high in the left side of the abdominal cavity, is a large, ovoid secondary lymphoid organ that plays a major role in mounting immune responses to antigens in the bloodstream (Figure 2-10). Whereas lymph nodes are specialized for encounters between lymphocytes and antigen drained from local tissues, the spleen specializes in filtering blood and trapping blood-borne antigens; thus, it is particularly important in the response to systemic infections. Unlike the lymph nodes, the spleen is not supplied by lymphatic vessels. Instead, blood-borne antigens and lymphocytes are carried into the spleen through the splenic artery and out via the splenic vein. Experiments with radioactively labeled lymphocytes show that more recirculating lymphocytes pass daily through the spleen than through all the lymph nodes combined. The spleen is surrounded by a capsule from which a number of projections (trabeculae) extend, providing structural support. Two main microenvironmental compartments can be distinguished in splenic tissue: the red pulp and white pulp, which are separated by a specialized region called the marginal zone (see Figure 2-10d). The splenic red pulp consists of a network of sinusoids populated by red blood cells, macrophages, and some lymphocytes. It is the site where old and defective red blood cells are destroyed and removed; many of the macrophages within the red pulp contain engulfed red blood cells or iron-containing pigments from degraded hemoglobin. It is also the site where pathogens first gain access to the lymphoid-rich regions of the spleen, known as the white pulp. The splenic white pulp surrounds the branches of the splenic artery, and consists of the periarteriolar lymphoid sheath (PALS) populated by T lymphocytes as well as B-cell follicles. As in lymph nodes, germinal centers are generated within these follicles during an immune response. The marginal zone, which borders the white pulp, is populated by unique and specialized macrophages and B cells, which are the first line of defense against certain blood-borne pathogens. Blood-borne antigens and lymphocytes enter the spleen through the splenic artery, and interact first with cells at the | CHAPTER 2 53 marginal zone. In the marginal zone, antigen is trapped and processed by dendritic cells, which travel to the PALS. Specialized, resident marginal zone B cells also bind antigen via complement receptors and convey it to the follicles. Migrating B and T lymphocytes in the blood enter sinuses in the marginal zone and migrate to the follicles and the PALS, respectively. The events that initiate the adaptive immune response in the spleen are analogous to those that occur in the lymph node. Briefly, circulating naïve B cells encounter antigen in the follicles, and circulating naïve CD8 and CD4 T cells meet antigen as MHC-peptide complexes on the surface of dendritic cells in the T-cell zone (PALS). Once activated, CD4 TH cells then provide help to B cells and CD8 T cells that have also encountered antigen. Some activated B cells, together with some TH cells, migrate back into follicles and generate germinal centers. It is unclear whether a reticular network operates as prominently within the spleen as it does in the lymph node. However, given that T cells, dendritic cells, and B cells find a way to interact efficiently within the spleen to initiate an immune response, it would not be surprising if a similar conduit system were present. Although animals can lead a relatively healthy life without a spleen, its loss does have consequences. In children, in particular, splenectomy (the surgical removal of a spleen) can lead to overwhelming post-splenectomy infection (OPSI) syndrome characterized by systemic bacterial infections (sepsis) caused by primarily Streptococcus pneumoniae, Neisseria meningitidis, and Haemophilus influenzae. Although fewer adverse effects are experienced by adults, splenectomy can still lead to an increased vulnerability to blood-borne bacterial infections, underscoring the role the spleen plays in our immune response to pathogens that enter the circulation. It is also important to recognize that the spleen has other functions (e.g., in iron metabolism, thrombocyte storage, hematopoiesis) that will also be compromised if it is removed. MALT Organizes the Response to Antigen That Enters Mucosal Tissues Lymph nodes and the spleen are not the only organs that develop secondary lymphoid microenvironments. T- and B-cell zones and lymphoid follicles are also found in mucosal membranes that line the digestive, respiratory, and urogenital systems, as well as in the skin. Mucosal membranes have a combined surface area of about 400 m2 (nearly the size of a basketball court) and are the major sites of entry for most pathogens. These vulnerable membrane surfaces are defended by a group of organized lymphoid tissues known collectively as mucosa-associated lymphoid tissue (MALT). Lymphoid tissue associated with different mucosal areas is sometimes given more specific names; for instance, the respiratory epithelium is referred to as bronchus-associated lymphoid tissue (BALT) or nasal-associated lymphoid tissue (NALT), and that associated with the intestinal epithelium is referred to as gut-associated lymphoid tissue (GALT). 54 PA R T I | Introduction (b) (a) Adenoids Tonsil Left subclavian vein Right lymphatic duct Lymph nodes Right subclavian vein Thymus Thoracic duct Spleen Peyer's patches Bone marrow Large intestine (c) White pulp Red pulp Small intestine Appendix Tissue lymphatics (d) Splenic artery Splenic vein Follicle Central arteriole Marginal zone White pulp Sinuses PALS Trabeculae Red pulp FIGURE 2-10 Structure of the spleen. (a) The spleen, which is about 5 inches long in human adults, is the largest secondary lymphoid organ. It is specialized for trapping blood-borne antigens. Panels (b) and (c) are stained tissue sections of the human spleen, showing the red pulp, white pulp, and follicles. These microenvironments are diagrammed schematically in (d). The splenic artery pierces the capsule and divides into progressively smaller arterioles, ending in vascular sinusoids that drain back into the splenic vein. The erythrocyte-filled red pulp surrounds the sinusoids. The white pulp forms a sleeve—the periarteriolar lymphoid sheath (PALS)— around the arterioles; this sheath contains numerous T cells. Closely associated with the PALS are the B-cell-rich lymphoid follicles that can develop into secondary follicles containing germinal centers. The marginal zone, a site of specialized macrophages and B cells, surrounds the PALS and separates it from the red pulp. [2-10b: Dr. Keith The structure of GALT is well described and ranges from loose, barely organized clusters of lymphoid cells in the lamina propria of intestinal villi to well-organized structures such as the tonsils and adenoids (Waldeyer’s tonsil ring), the appendix, and Peyer’s patches, which are found within the intestinal lining and contain well-defined follicles and T-cell zones. As shown in Figure 2-11, lymphoid cells are found in various regions within the lining of the intestine. The outer mucosal epithelial layer contains intraepithelial lymphocytes (IELs), many of which are T cells. The lamina propria, which lies under the epithelial Wheeler/Photo Researchers; 2-10c: Biophoto Associates/Photo Researchers.] Cells, Organs, and Microenvironments of the Immune System | CHAPTER 2 55 (b) (a) MALT Adenoids Tonsil Thoracic duct Thymus Right lymphatic duct Right subclavian vein Left subclavian vein Lymph nodes Spleen Thoracic duct Peyer's patches Large intestine (c) Bone marrow Villi Small intestine Appendix M cells Lamina propria Submucosa High endothelial venule Tissue lymphatics Lymphocytes Lymph vessel Lymphocytes B cells Germinal center T cells Muscle layer Peyer’s patch FIGURE 2-11 Mucosa-associated lymphoid tissue (MALT). (a) The Peyer’s patch is a representative of the extensive MALT system that is found in the intestine. (b) A stained tissue cross-section of Peyer’s patch lymphoid nodules in the intestinal submucosa is schematically diagrammed in (c). The intestinal lamina propria contains loose clusters of lymphoid cells and diffuse follicles. [2-11b: Dr. Gladden Willis/Visuals Unlimited.] layer, contains large numbers of B cells, plasma cells, activated T cells, and macrophages in loose clusters. Microscopy has revealed more than 15,000 lymphoid follicles within the intestinal lamina propria of a healthy child. Peyer’s patches, nodules of 30 to 40 lymphoid follicles, extend into the muscle layers that are just below the lamina propria. Like lymphoid follicles in other sites, those that compose Peyer’s patches can develop into secondary follicles with germinal centers. The overall functional importance of MALT in the body’s defense is underscored by its large population of antibody-producing plasma cells, whose number exceeds that of plasma cells in the spleen, lymph nodes, and bone marrow combined. Some cellular structures and activities are unique to MALT. For instance, the epithelial cells of mucous membranes play an important role in delivering small samples of foreign antigen from the respiratory, digestive, and urogenital tracts to the underlying mucosa-associated lymphoid tissue. In the digestive tract, specialized M cells transport antigen across the epithelium (Figure 2-12). The structure of M cells is striking: they are flattened epithelial cells lacking the microvilli that characterize the rest of the mucosal epithelium. They have a deep invagination, or pocket, in the basolateral plasma membrane, which is filled with a cluster of B cells, T cells, and macrophages. Antigens in the intestinal lumen are endocytosed into vesicles that are transported from the luminal membrane to the underlying pocket membrane. The vesicles then fuse with the pocket membrane, delivering antigens to clusters of lymphocytes and antigen-presenting cells, the most important of which are dendritic cells, contained within the pocket. Antigen transported across the mucous membrane by M cells ultimately leads to the activation of B cells that differentiate and then secrete IgA. This class of antibody is concentrated in 56 PA R T I | Introduction (a) (b) M cell Lumen Antigen Antigen TH cell Mucosal epithelium Intraepithelial lymphocyte IgA M cell IgA Pocket Lamina propria B cells Plasma cell Organized lymphoid follicle Macrophage (c) M cell Macrophage FIGURE 2-12 Structure of M cells and production of IgA at inductive sites. (a) M cells, situated in mucous membranes, endocytose antigen from the lumen of the digestive, respiratory, and urogenital tracts. The antigen is transported across the cell and released into the large basolateral pocket. (b) Antigen transported across the epithelial layer by M cells at an inductive site activates B cells in the underlying lymphoid follicles. The activated B cells differentiate into IgA-producing plasma cells, which migrate along the lamina propria, the layer under the mucosa. secretions (e.g., milk) and is an important tool used by the body to combat many types of infection at mucosal sites. The Skin Is an Innate Immune Barrier and Also Includes Lymphoid Tissue The skin is the largest organ in the body and a critical anatomic barrier against pathogens. It also plays an important role in nonspecific (innate) defenses (Figure 2-13). The epidermal (outer) layer of the skin is composed largely of specialized epithelial cells called keratinocytes. These cells secrete a number of cytokines that may function to induce a local inflammatory The outer mucosal epithelial layer contains intraepithelial lymphocytes, of which many are T cells. (c) A stained section of mucosal lymphoid tissue (the Peyer’s patch of the intestine) shows small, darkly stained intraepithelial lymphocytes encased by M cells, whose nuclei are labeled. Lymphocytes are also present in the lamina propria. [2-12c: Kucharzik, T., Lugering, N., Schmid, K.W., Schmidt, M.A., Stoll, R. Domschke, W. Human intestinal M cells exhibit enterocyte-like intermediate filaments. Gut 1998, 42: 54–52. doi: 10.1136/gut.42.1.54.] reaction. Scattered among the epithelial-cell matrix of the epidermis are Langerhans cells, skin-resident dendritic cells that internalize antigen by phagocytosis or endocytosis. These Langerhans cells undergo maturation and migrate from the epidermis to regional lymph nodes, where they function as potent activators of naïve T cells. In addition to Langerhans cells, the epidermis also contains intraepidermal lymphocytes, which are predominantly T cells; some immunologists believe that they play a role in combating infections that enter through the skin, a function for which they are well positioned. The underlying dermal layer of the skin also contains scattered lymphocytes, dendritic cells, monocytes, macrophages, and Cells, Organs, and Microenvironments of the Immune System Epidermis Keratinocyte Langerhans cell | CHAPTER 2 57 may even include hematopoietic stem cells. Most skin lymphocytes appear to be either previously activated cells or memory cells, many of which traffic to and from local, draining lymph nodes that coordinate the responses to pathogens that have breached the skin barrier. Dermis Intraepidermal lymphocyte Blood vessel Tertiary Lymphoid Tissues Also Organize and Maintain an Immune Response Dendritic cell Tissues that are the sites of infection are referred to as tertiary lymphoid tissue. Lymphocytes activated by antigen in secondary lymphoid tissue can return to these organs (e.g., lung, liver, brain) as effector cells and can also reside there as memory cells. It also appears as if tertiary lymphoid tissues can generate defined microenvironments that organize the returning lymphoid cells. Investigators have recently found that the brain, for instance, establishes reticular systems that guide lymphocytes responding to chronic infection with the protozoan that causes toxoplasmosis. These observations together highlight the remarkable adaptability of the immune system, as well as the intimate relationship between anatomical structure and immune function. The conservation of structure/function relationships is also illustrated by the evolutionary relationships among immune systems and organs (see the Evolution Box 2-4 below on pages 57–59). Lymphatic vessel FIGURE 2-13 The distribution of immune cells in the skin. Langerhans cells reside in the outer layer, the epidermis. They travel to lymph nodes via the lymphatic vessels in the dermis, a layer of connective tissue below the epidermis. Dendritic cells also reside in the dermis and also can travel via the lymphatic vessels to lymph nodes when activated. White blood cells, including monocytes and lymphocytes travel to both layers of the skin via blood vessels, extravasating in the dermis, as shown. BOX 2-4 EVOLUTION Variations on Anatomical Themes In Chapter 5, we will see that innate systems of immunity are found in vertebrates, invertebrates, and even in plants. Adaptive immunity, which depends on lymphocytes and is mediated by antibodies and T cells, only evolved in the subphylum Vertebrata. However, as shown in Figure 1, the kinds of lymphoid tissues seen in different orders of vertebrates differ dramatically. All multicellular living creatures, plant and animal, have cellular and molecular systems of defense against pathogens. Vertebrates share an elaborate anatomical immune system that is compartmentalized in several predictable ways. For example, sites where immune cells develop (primary lymphoid organs) are separated from where they generate an immune response (secondary lymphoid tissue). Likewise, sites that support the development of B cells and myeloid cells, are separated from sites that generate mature T cells. The spectrum of vertebrates ranges from the jawless fishes (Agnatha, the earliest lineages, which are represented by the lamprey eel and hagfish), to cartilaginous fish (sharks, rays), which represent the earliest lineages of jawed vertebrates (Gnathosomata), to bony fish, reptiles, amphibians, birds, and mammals, the most recently evolved vertebrates. If you view these groups as part of an evolutionary progression, you see that, in general, immune tissues and organs evolved by earlier orders have been retained as newer organs of immunity, such as lymph nodes, have appeared. For example, all vertebrates have gut-associated lymphoid tissue (GALT), but only jawed vertebrates have a well-developed thymus and spleen. The adaptive immune system of jawed vertebrates emerged about 500 million years ago. T cells were the first cell population to express a diverse repertoire of anti- gen receptors, and their appearance is directly and inextricably linked to the appearance of the thymus. This dependence is reflected in organisms today: all jawed vertebrates have a thymus, and the thymus is absolutely required for the development of T cells. Until recently, it was thought that jawless vertebrates (e.g., the lamprey eel) did not have any thymic tissue. However, recent studies indicate that even the lamprey supports the development of two distinct lymphocyte populations reminiscent of T and B cells, and may also harbor distinct thymic-like tissue in their gill regions (Figure 2). In contrast, their B-lymphocytelike cells may develop in distinct regions associated with the kidneys and intestine (typhlosole). These findings suggest that jawed and jawless vertebrates share a common ancestor in which lymphocyte lineages and primary lymphoid organs were already compartmentalized. (continued) 58 PA R T I EVOLUTION | Introduction (continued) Thymus Lymph nodes Thymus Thymus Kidney GALT Thymus GALT GALT Peyer's patch GALT Spleen Spleen Bone marrow Lymph nodes Bone marrow Spleen Lamprey GALT Spleen Trout Frog Bone marrow Bursa Chicken Mouse GALT Thymus Spleen Bone marrow Lymph nodes Germinal centers ? Anura Teleostei Aves Mammalia Reptilia Amphibia Osteichthyes Agnatha Gnathostomata Vertebrata FIGURE 1 Evolutionary distribution of lymphoid tissues. The presence and location of lymphoid tissues in several major orders of vertebrates are shown. Although they are not shown in the diagram, cartilaginous fish such as sharks and rays have GALT, a thymus, and a spleen. Reptiles also have GALT, a thymus, and a spleen and may also have lymph nodes that participate in immunological reactions. The sites and nature of primary lymphoid tissues in reptiles are under investigation. Although jawless animals were thought not to have a thymus, recent data suggest that they might have thymic tissue and two types of lymphocytes analogous to B cells and T cells. [Adapted from Dupasquier and M. Flajnik, 2004, in Fundamental Immunology, 5th ed., W. E. Paul, ed., Lippincott-Raven, Philadelphia.] Whereas T-cell development is inextricably linked to the presence of a thymus, B-cell development does not appear to be bound to one particular organ. Although the bone marrow is the site of B-cell development in many mammals, including mouse and human, it is not the site of B-cell development in all species. For instance, the anatomical sites of B-cell development shift over the course of development in sharks (from liver to kidney to spleen), in amphibians and reptiles (from liver and spleen to bone marrow). In contrast, B-cell development in bony fish takes place primarily within the kidney. In birds, B cells complete their development in a lymphoid organ associated with the gut, the bursa of Fabricius (Figure 3). The gut is also a site of B-cell development in some mammals. In cattle and sheep, early in gestation, B cells mature in the fetal spleen. Later in gestation, however, this function is assumed by the Cells, Organs, and Microenvironments of the Immune System | CHAPTER 2 59 BOX 2-4 (c) (b) (a) FIGURE 2 Thymic tissue in the lamprey eel. (a) Jawless vertebrates, including the lamprey eel (shown here in larval form), were thought not to have a thymus. Recent work suggests, however, that they generate two types of lymphocytes analogous to B and T cells and have thymic tissue at the tip of their gills (b). The thymic tissue shown in (c) is stained blue for a gene (CDA1) specific for lymphocytes found in lampreys. The thymus of jawed vertebrates arises from an area analogous to the gill region. [Left: Courtesy of Brian Morland, The Bellflask Ecological Survey Team; middle: Dr. Keith Wheeler/Science Photo Library/ Photo Researchers; right: Courtesy Thomas Boehm.] patch of tissue embedded in the wall of the intestine called the ileal Peyer’s patch, which contains a large number of B cells as well as T cells. The rabbit, too, uses gutassociated tissues, especially the appendix, as primary lymphoid tissue for important steps in the proliferation and diversification of B cells. Recent work suggests that even in animals that depend largely on the bone marrow to complete B cell maturation, lymphoid tissue in the gut can act as a primary lymphoid organ for generating mature B lymphocytes. Secondary lymphoid organs are more variable in their numbers and location than primary lymphoid organs (e.g., rodents do not have tonsils, but do have well-developed lymphoid tissue at the base of the nose). The structural and functional features of these organs are, however, shared by most vertebrates with at least one interesting exception: pig lymph nodes exhibit a striking peculiarity—they are “inverted” anatomically, so that the medulla of the organ, where lymphocytes exit the organ, is on the outside, and the cortex, where lymphocytes meet their antigen and proliferate, is on the inside. Lymphocyte egress is also inverted: they exit via blood vessels, rather than via efferent lymphatics. The adaptive advantages (if any) of these odd structural variations are unknown, but the example reminds us of the remarkable plasticity of structure/function relationships within biological structures and the creative opportunism of evolutionary processes. (b) (a) Bursa of Fabricius (B cell development) Thymus (T cell development) Intestines FIGURE 3 The avian bursa. (a) Like all vertebrates, birds have a thymus, spleen, and lymph nodes, and hematopoiesis occurs in their bone marrow. However, their B cells do not develop in the bone marrow. Rather they develop in an outpouching of the intestine, the bursa, which is located close to the cloaca (the common end of the intestinal and genital tracts in birds). A stained tissue section of the bursa and cloaca is shown in (b). [(b) Courtesy Dr. Thomas Caceci.] 60 PA R T I | Introduction S U M M A R Y ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ The immune response results from the coordinated activities of many types of cells, organs, and microenvironments found throughout the body. Many of the body’s cells, tissues, and organs arise from different stem cell populations. All red and white blood cells develop from a pluripotent hematopoietic stem cell during a highly regulated process called hematopoiesis. In the adult vertebrate, hematopoiesis occurs in the bone marrow, a stem cell niche that supports both the selfrenewal of stem cells and their differentiation into multiple blood cell types. Hematopoietic stem cells give rise to two main blood cell progenitors: common myeloid progenitors and common lymphoid progenitors. Four main types of cells develop from common myeloid progenitors: red blood cells (erythrocytes), monocytes (which give rise to macrophages and myeloid dendritic cells), granulocytes (which include the abundant neutrophils and less abundant basophils, eosinophils, and mast cells), and megakaryocytes. Macrophages and neutrophils are specialized for the phagocytosis and degradation of antigens. Macrophages also have the capacity to present antigen with MHC to T cells. Immature forms of dendritic cells have the capacity to capture antigen in one location, undergo maturation, and migrate to another location, where they present antigen to T cells. Dendritic cells are the most potent antigen-presenting cell for activating naïve T cells. There are three types of lymphoid cells: B cells, T cells, and natural killer (NK) cells. B and T cells are members of clonal populations distinguished by antigen receptors of unique specificity. B cells synthesize and display membrane antibody, and T cells synthesize and display T-cell receptors (TCRs). NK cells do not synthesize antigenspecific receptors; however a small population of TCRexpressing T cells have features of NK cells and are called NKT cells. T cells can be further subdivided into helper T cells, which typically express CD4 and see peptide bound to MHC class II, and cytotoxic T cells, which typically express CD8 and see peptide bound to MHC class I. Primary lymphoid organs are the sites where lymphocytes develop and mature. T cell precursors come from the bone marrow but develop fully in the thymus; in humans and mice, B cells arise and develop in bone marrow. In birds, B cells develop in the bursa. ■ ■ ■ ■ ■ ■ ■ ■ Secondary lymphoid organs provide sites where lymphocytes encounter antigen, become activated, and undergo clonal expansion and differentiation into effector cells. They include the lymph node, spleen, and more loosely organized sites distributed throughout the mucosal system (the mucosa-associated lymphoid tissue, or MALT) and are typically compartmentalized into T-cell areas and B-cell areas (follicles). Lymph nodes and the spleen are the most highly organized secondary lymphoid organs. T-cell and B-cell activity are separated into distinct microenvironments in both. T cells are found in the paracortex of the lymph nodes and the periarteriolar sheath of the spleen. B cells are organized into follicles in both organs. The spleen is the first line of defense against blood-borne pathogens. It also contains red blood cells and compartmentalizes them in the red pulp from lymphocytes in the white pulp. A specialized region of macrophages and B cells, the marginal zone, borders the white pulp. In secondary lymphoid tissue, T lymphocytes can develop into killer and helper effector cells. CD4 T cells can help B cells to differentiate into plasma cells, which secrete antibody or can also develop into effector cells that further activate macrophages and CD8 cytotoxic T cells. Both B and T cells also develop into long-lived memory cells. B cells undergo further maturation in germinal cells, a specialized substructure found in follicles after infection; here they can increase their affinity for antigen and undergo class switching. Lymphoid follicles and other organized lymphoid microenvironments are found associated with the mucosa, the skin, and even tertiary tissues at the site of infection. Mucosa-associated lymphoid tissue (MALT) is an important defense against infection at epithelial layers and includes a well-developed network of follicles and lymphoid microenvironments associated with the intestine (gut-associated lymphoid tissue, GALT). Unique cells found in the lining of the gut, M cells, are specialized to deliver antigen from the intestinal spaces to the lymphoid cells within the gut wall. R E F E R E N C E S Banchereau J., et al. 2000. Immunobiology of dendritic cells. Annual Review of Immunology 18:767. Directing lymphocyte traffic in the lymph node. Trends in Immunology 28:346–352. Bajenoff, M., J. G. Egen, H. Qi, A. Y. C. Huang, F. Castellino, and R. N. Germain. 2007. Highways, byways and breadcrumbs: Bajoghil, B., P. Guo, N. Aghaallaei, M. Hirano, C. Strohmeier, N. McCurley, D. E. Bockman, M. Schorpp, M. D. Cooper, and Cells, Organs, and Microenvironments of the Immune System T. Boehm. 2011. A thymus candidate in lampreys. Nature 47:90–95. Boehm, T., and C. E. Bleul. 2007. The evolutionary history of lymphoid organs. Nature Immunology 8:131–135. Borregaard, N. 2011. Neutrophils: From marrow to microbes. Immunity 33:657–670. Catron, D.M., A. A. Itano, K. A. Pape, D. L. Mueller, and M. K Jenkins. 2004. Visualizing the first 50 hr of the primary immune response to a soluble antigen. Immunity 21:341–347. Ema, H., et al. 2005. Quantification of self-renewal capacity in single hematopoietic stem cells from normal and Lnk-deficient mice. Developmental Cell 6:907. | CHAPTER 2 61 Piccirillo, C. A., and E. M. Shevach. 2004. Naturally occurring CD4CD25 immunoregulatory T cells: Central players in the arena of peripheral tolerance. Seminars in Immunology 16:81. Picker, L. J., and M. H. Siegelman. 1999. Lymphoid tissues and organs. In Fundamental Immunology, 4th ed. W. E. Paul, ed. Philadelphia: Lippincott-Raven, p. 145. Rodewald, H-R. 2006. A second chance for the thymus. Nature 441:942–943. Rothenberg, M., and S. P. Hogan. 2006. The eosinophil. Annual Review of Immunology 24:147–174. Scadden, D. T. 2006. The stem-cell niche as an entity of action. Nature 441:1075–1079. Falcone, F. H., D. Zillikens, and B. F.Gibbs. 2006. The 21st century renaissance of the basophil? Current insights into its role in allergic responses and innate immunity. Experimental Dermatology 15:855–864. Shizuru, J. A., R. S. Negrin, and I. L. Weissman. 2005. Hematopoietic stem and progenitor cells: Clinical and preclinical regeneration of the hematolymphoid system. Annual Review of Medicine 56:509. Godfrey, D. I., H. R. MacDonald, M. Kronenberg, M. J. Smyth, and L. Van Kaer. 2004. NKT cells: What’s in a name? Nature Reviews Immunology 4:231. Weissman, I. L. 2000. Translating stem and progenitor cell biology to the clinic: Barriers and opportunities. Science 287:1442. Halin, C., J. R. Mora, C. Sumen, and U. H. von Andrian. In vivo imaging of lymphocyte trafficking. 2005 Annual Review of Cell and Developmental Biology 21:581–603. Useful Web Sites Kiel, M., and S. J. Morrison. 2006. Maintaining hematopoietic stem cells in the vascular niche. Immunity 25:852–854. Liu, Y. J. 2001. Dendritic cell subsets and lineages, and their functions in innate and adaptive immunity. Cell 106:259. Mebius, R. E., and G. Kraal. 2005. Structure and function of the spleen. Nature Reviews Immunology. 5:606–616. Mendez-Ferrer, S., D. Lucas, M. Battista, and P. S. Frenette. 2008. Haematopoietic stem cell release is regulated by circadian oscillations. Nature 452:442–448. http://bio-alive.com/animations/anatomy.htm A collection of publicly available animations relevant to biology. Scroll through the list to find videos on blood and immune cells, immune responses, and links to interactive sites that reinforce your understanding of immune anatomy. http://stemcells.nih.gov Links to information on stem cells, including basic and clinical information, and registries of available embryonic stem cells. www.niaid.nih.gov/topics/immuneSystem/Pages/ structureImages.aspx A very accessible site about Mikkola, H. K. A., and S. A. Orkin. 2006. The journey of developing hematopoietic stem cells. Development 133:3733–3744. immune system function and structure. Miller, J. F. 1961. Immunological function of the thymus. Lancet 2:748–749. www.immunity.com/cgi/content/full/21/3/341/ DC1 A pair of simulations that trace the activities of T cells, Miller, J. F. A. P. 2002. The discovery of thymus function and of thymus-derived lymphocytes. Immunological Reviews 185:7–14. B cells, and dendritic cells in a lymph node. These movies are discussed in more detail in Chapter 14. Moon, J. J., H. H. Chu, M. Pepper, S. J. McSorley, S. C. Jameson, R. M. Kedl, and M. K. Jenkins. 2007. Naïve CD4 T cell frequency varies for different epitopes and predicts repertoire diversity and response magnitude. Immunity 27:203–213. www.hhmi.org/research/investigators/cyster. html Investigator Jason Cyster’s public website that includes Pabst, R. 2007. Plasticity and heterogeneity of lymphoid organs: What are the criteria to call a lymphoid organ primary, secondary or tertiary? Immunology Letters 112:1–8. S T U D Y many videos of B-cell activity in immune tissue. www.hematologyatlas.com/principalpage.htm An interactive atlas of both normal and pathological human blood cells. Q U E S T I O N S RESEARCH FOCUS QUESTION Notch is a surface protein that regulates cell fate. When bound by its ligand, it releases and activates its intracellular region, which regulates new gene transcription. Investigators found that the phenotype of de- veloping cells in the bone marrow differed dramatically when they overexpressed the active, intracellular portion of Notch. In particular, the frequency of BCR cells plummeted, and the frequency of TCR cells increased markedly. Interestingly, 62 PA R T I | Introduction other investigators found that when you knocked out Notch, the phenotype of cells in the thymus changed: the frequency of BCR cells increased and the frequency of TCR cells decreased dramatically. Propose a molecular model to explain these observations, and an experimental approach to begin testing your model. c. TC cells and basophils d. TH and B cells 3. List two primary and two secondary lymphoid organs and summarize their functions in the immune response. 4. What two primary characteristics distinguish hematopoi- etic stem cells from mature blood cells? CLINICAL FOCUS QUESTION The T and B cells that differenti- ate from hematopoietic stem cells recognize as self the bodies in which they differentiate. Suppose a woman donates HSCs to a genetically unrelated man whose hematopoietic system was totally destroyed by a combination of radiation and chemotherapy. Further, suppose that, although most of the donor HSCs differentiate into hematopoietic cells, some differentiate into cells of the pancreas, liver, and heart. Decide which of the following outcomes is likely and justify your choice. a. The T cells that arise from the donor HSCs do not attack the pancreatic, heart, and liver cells that arose from donor cells but mount a GVH response against all of the other host cells. b. The T cells that arise from the donor HSCs mount a GVH response against all of the host cells. c. The T cells that arise from the donor HSCs attack the pancreatic, heart, and liver cells that arose from donor cells but fail to mount a GVH response against all of the other host cells. d. The T cells that arise from the donor HSCs do not attack the pancreatic, heart, and liver cells that arose from donor cells and fail to mount a GVH response against all of the other host cells. 1. Explain why each of the following statements is false. a. There are no mature T cells in the bone marrow. b. The pluripotent stem cell is one of the most abundant cell types in the bone marrow. c. There are no stem cells in blood. d. Activation of macrophages increases their expression of class I MHC molecules, allowing them to present antigen to TH cells more effectively. e. Mature B cells are closely associated with osteoblasts in the bone marrow. f. Lymphoid follicles are present only in the spleen and lymph nodes. g. The FRC guides B cells to follicles. h. Infection has no influence on the rate of hematopoiesis. i. Follicular dendritic cells can process and present antigen to T lymphocytes. j. Dendritic cells arise only from the myeloid lineage. k. All lymphoid cells have antigen-specific receptors on their membrane. l. All vertebrates generate B lymphocytes in bone marrow. m. All vertebrates have a thymus. n. Jawless vertebrates do not have lymphocytes. 2. For each of the following sets of cells, state the closest com- mon progenitor cell that gives rise to both cell types. a. Dendritic cells and macrophages b. Monocytes and neutrophils 5. What are the two primary roles of the thymus? 6. At what age does the thymus reach its maximal size? a. During the first year of life b. Teenage years (puberty) c. Between 40 and 50 years of age d. After 70 years of age 7. Preparations enriched in hematopoietic stem cells are useful for research and clinical practice. What is the role of the SCID mouse in demonstrating the success of HSC enrichment? 8. Explain the difference between a monocyte and a macrophage. 9. What effect would removal of the bursa of Fabricius (bur- sectomy) have on chickens? 10. Indicate whether each of the following statements about the lymph node and spleen is true or false. If you think a statement is false, explain why. a. The lymph node filters antigens out of the blood. b. The paracortex is rich in T cells, and the periarteriolar lymphoid sheath (PALS) is rich in B cells. c. Only the lymph node contains germinal centers. d. The FRCC enhances T cell/APC interactions e. Afferent lymphatic vessels draining the tissue spaces enter the spleen. f. Lymph node but not spleen function is affected by a knockout of the Ikaros gene. 11. For each description below (1–14), select the appropriate cell type (a–p). Each cell type may be used once, more than once, or not at all. Descriptions 1. Major cell type presenting antigen to naïve T cells 2. Phagocytic cell of the central nervous system 3. Granulocytic cells important in the body’s defense against parasitic organisms 4. Give rise to red blood cells 5. Generally first cells to arrive at site of inflammation 6. Support maintenance of hematopoietic stem cells 7. Give rise to thymocytes 8. Circulating blood cells that differentiate into macro- phages in the tissues 9. An antigen-presenting cell that arises from the same precursor as a T cell but not the same as a macrophage 10. Cells that are important in sampling antigens of the intestinal lumen Cells, Organs, and Microenvironments of the Immune System 11. Granulocytic cells that release various pharmacologically active substances 12. White blood cells that play an important role in the devel- opment of allergies 13. Cells that can use antibodies to recognize their targets 14. Cells that express antigen-specific receptors Cell Types a. Common myeloid progenitor cells b. Monocytes c. Eosinophils | CHAPTER 2 d. Dendritic cells e. Natural killer (NK) cells f. Mast cells g. Neutrophils h. M cells i. Osteoblasts j. Lymphocytes k. NK1-T cell l. Microglial cell m. Myeloid dendritic cell n. Hematopoietic stem cell o. Lymphoid dendritic cell 63 This page lelt intentionally blank. 3 Receptors and Signaling: B and T-Cell Receptors T he coordination of physiological functions throughout the body depends on the ability of individual cells to sense changes in their environment and to respond appropriately. One of the major routes by which a cell interprets its surroundings is through the binding of signaling molecules to cell-associated receptor proteins. A molecule that binds to a receptor is a ligand. Noncovalent binding of a ligand to its receptor may induce alterations in the receptor itself, in its polymerization state, and/or in the environment of that receptor. These changes act to transmit or transduce the ligand-binding signal into the interior of the cell, leading to alterations in cellular functions. In the nervous system, we would call the signaling molecules neurotransmitters; in the endocrine system, hormones. In the immune system, the foreign molecules that signal the presence of non-self entities are antigens, and the small molecules that communicate among the various populations of immune cells are cytokines. Specialized cytokines that induce chemo-attraction or -repulsion are termed chemokines. In this chapter, we provide a general introduction to receptor-ligand binding and to the broad concepts and strategies that underlie signal transduction. We then focus specifically on the antigens and receptors of the adaptive immune system, introducing the B- and T-cell receptors and the intracellular signaling events that occur upon antigen binding. Because the cells of the immune system are distributed throughout the body—with some resident in fixed tissues and others circulating through the various lymphoid tissues, the blood, and the lymphatics—the ability of these cells to communicate with and signal to one another via soluble cytokine and chemokine molecular messengers is essential to their function. In Chapter 4, we describe the signaling events that result when cytokines and chemokines bind to their cognate (matching) receptors. Chapter 5 includes a description of the pattern recognition receptors (PRRs) of the innate immune system and of the signaling events initiated by their antigen binding. All signaling events begin with a complementary interaction between a ligand and a receptor. This figure depicts the molecular interaction between the variable regions of an antibody molecule (light and heavy chains are shown in blue and red, respectively) and the tip of the influenza hemagglutinin molecule, shown in yellow. [Illustration based on x-ray crystallography data collected by P. M. Colman and W. R. Tulip from GJVH Nossal, 1993, Scientific American 269(3):22.] ■ ■ Receptor-Ligand Interactions Common Strategies Used in Many Signaling Pathways ■ Frequently Encountered Signaling Pathways ■ The Structure of Antibodies ■ Signal Transduction in B Cells ■ T-Cell Receptors and Signaling The essential concepts that underlie cell signaling in the immune system can be summarized as follows: 65 • A cellular signal is any event that instructs a cell to change its metabolic or proliferative state. • Signals are usually generated by the binding of a ligand to a complementary cell-bound receptor. • A cell can become more or less susceptible to the actions of a ligand by increasing (up-regulating) or decreasing (downregulating) the expression of the receptor for that ligand. 66 PA R T I | Introduction • The ligand may be a soluble molecule, or it may be a peptide, carbohydrate, or lipid presented on the surface of a cell. • The ligand may travel long distances from its entry point through the body in either the bloodstream or the lymphatics before it reaches a cell bearing the relevant receptor. • Ligand-receptor binding is noncovalent, although it may be of quite high affinity. • Ligand binding to the receptor induces a molecular change in the receptor. This change may be in the form of a conformational alteration in the receptor, receptor dimerization or clustering, a change in the receptor’s location in the membrane, or a covalent modification. • Such receptor alterations set off cascades of intracellular events that include the activation of enzymes, and changes in the intracellular locations of important molecules. • The end result of cellular signaling is often, but not always, a change in the transcriptional program of the target cell. • Sometimes a cell must receive more than one signal through more than one receptor in order to effect a particular outcome. • Integration of all the signals received by a cell occurs at the molecular level inside the recipient cell. are individually weak, many such interactions are required to form a biologically significant receptor-ligand connection. Furthermore, since each of these noncovalent interactions operates only over a very short distance—generally about 1 Angstrom, (1 Å  10⫺10 meters)—a high-affinity receptorligand interaction depends on a very close “fit,” or degree of complementarity, between the receptor and the ligand (Figure 3-1). How Do We Quantitate the Strength of Receptor-ligand Interactions? Consider a receptor, R, binding to a ligand, L. We can describe their binding reaction according to the following equation, k1 R ⫹ L ∆ RL k⫺1 (Eq. 3-1) in which RL represents the bound receptor-ligand complex, k1 is the forward or association rate constant, and k⫺1 is the reverse or dissociation rate constant. The ratio of k1/k⫺1 is equal to Ka, the association constant of the reaction, and is a measure of the affinity of the Ligand Receptor NH2 Receptor-Ligand Interactions The antigen receptors of the adaptive immune system are transmembrane proteins localized at the plasma membrane. Ligand binding to its cognate receptor normally occurs via specific, noncovalent interactions between the ligand and the extracellular portion of the membrane receptor. Although a single lymphocyte expresses only one type of antigen receptor, it also may express many different receptor molecules for signals such as cytokines and chemokines, and therefore a healthy cell must integrate the signals from all the receptors that are occupied at any one time. CH2 OH ••• O CH2 CH2 NH3+ –O C CH2 C CH2 Hydrogen bond CH2 CH2 Ionic bond O CH2 CH CH3 Hydrophobic interactions CH3 CH3 CH3 CH CH CH2 van der Waals interactions CH3 CH3 CH CH CH3 O Receptor-Ligand Binding Occurs via Multiple Noncovalent Bonds The surface of a receptor molecule binds to its complementary ligand surface by the same types of noncovalent chemical linkages that enzymes use to bind to their substrates. These include hydrogen and ionic bonds, and hydrophobic and van der Waals interactions. The key to a meaningful receptor-ligand interaction is that the sum total of these bonding interactions hold the two interacting surfaces together with sufficient binding energy, and for sufficient time, to allow a signal to pass from the ligand to the cell bearing the receptor. Because these noncovalent interactions CH2 +H C O– 3N CH2 Ionic bond FIGURE 3-1 Receptor-ligand binding obeys the rules of chemistry. Receptors bind to ligands using the full-range of noncovalent bonding interactions, including ionic and hydrogen bonds and van der Waals and hydrophobic interactions. For signaling to occur, the bonds must be of sufficient strength to hold the ligand and receptor in close proximity long enough for downstream events to be initiated. In B- and T-cell signaling, activating interactions also require receptor clustering. In an aqueous environment, noncovalent interactions are weak and depend on close complementarity of the shapes of receptor and ligand. Receptors and Signaling: B and T-Cell Receptors receptor-ligand pair. The association constant is defined as the relationship between the concentration of reaction product, [RL], and the product of the concentrations of reactants, [R] multiplied by [L]. The units of affinity are therefore M⫺1. The higher the Ka, the higher the affinity of the interaction. Ka  3RL4 3R 4 3L4 (Eq. 3-2) The reciprocal of the association constant, the dissociation constant, Kd, is often used to describe the interactions between receptors and ligands. It is defined as: Kd  3R4 3L 4 3RL4 (Eq. 3-3) The units of the dissociation constant are in molarity (M). Inspection of this equation reveals that when half of the receptor sites are filled with ligand—that is, [R][RL]—the Kd is equal to the concentration of free ligand [L]. The lower the Kd, the higher the affinity of the interaction. For purposes of comparison, it is useful to consider that the Kd values of many enzyme-substrate interactions lie in the range of 10⫺3 M to 10⫺5 M. The Kd values of antigenantibody interactions at the beginning of an immune response are normally on the order of 10⫺4 to 10⫺5 M. However, since the antibodies generated upon antigen stimulation are mutated and selected over the course of an immune response, antigen-antibody interactions late in an immune response may achieve a Kd as low as 10⫺12 M. Under these conditions, if an antigen is present in solution at a concentration as low as 10⫺12 M, half of the available antibody-binding sites will be filled. This is an extraordinarily strong interaction between receptor and ligand. Interactions Between Receptors and Ligands Can Be Multivalent Many biological receptors, including B-cell receptors, are multivalent—that is, they have more than one ligand binding site per molecule. When both receptors and ligands are multivalent—as occurs, for example, when a bivalent immunoglobulin receptor on the surface of a B cell binds to two, identical, repeated antigens on a bacterial surface—the overall binding interaction between the bacterial cell and the B-cell receptor is markedly enhanced compared with a similar, but univalent, interaction. In this way, multiple concurrent receptor-ligand interactions increase the strength of binding between two cell surfaces. Note, however, that binding via two identical receptor sites to two identical ligands on the same cell may be a little less than twice as firm as binding through a single receptor site. This is because the binding of both receptor sites to two ligands on a single antigen may somewhat strain the geometry of the binding at one or both of the sites and therefore slightly interfere with the “fit” of the individual interactions. Much of the benefit of multivalency results from the fact that noncovalent-binding interactions are inherently | CHAPTER 3 67 reversible; the ligand spends some of its time binding to the receptor, and some of its time in an unbound, or “off ” state (Figure 3-2a). When more than one binding site is involved, it is less likely that all of the receptor sites will simultaneously be in the “off ” state, and therefore that the receptor will release the ligand (Figures 3-2b and 3-2c). The term avidity is used to describe the overall strength of the collective binding interactions that occur during multivalent binding. In the early phases of the adaptive immune response to multivalent antigens, B cells secrete IgM antibody, which has 10 available antigen binding sites per molecule. IgM is therefore capable of binding to antigens at a biologically significant level, even if the affinity of each individual binding site for its antigen is low, because of the avidity of the entire antigen-antibody interaction. Likewise, when a cell-bound receptor binds to a cell-bound ligand, their interaction may be functionally multivalent, even if the individual receptor molecules are monovalent, because multiple receptor molecules can cluster in the cell membrane and participate in the receptor-ligand interaction. In addition, other co-receptor interactions can also contribute to the overall avidity of the engagement between a cell and its antigen (see below). The affinity of receptor-ligand interactions can be measured by techniques such as equilibrium dialysis or surface plasmon resonance (SPR). Both of these methods are described in Chapter 20. (a) “On” “Off” FIGURE 3-2 Monovalent and bivalent binding. (a) Monovalent binding. The receptor exists in equilibrium with its ligand, represented here as a circle. Part of the time it is bound (binding is in the “on” state), and part of the time it is unbound (binding is in the “off ” state). The ratio of the time spent in the “on” versus the “off ” state determines the affinity of the receptor-ligand interaction and is related to the strength of the sum of the noncovalent binding interactions between the receptor and the ligand. PA R T I 68 | Introduction (b) (c) FIGURE 3-2 (continued) (b) Bivalent binding. Here, we visualize a bivalent receptor, binding to two sites on a single multivalent ligand. This could represent, for example, a bivalent antibody molecule binding to a bacterium with repeated antigens on its surface. In this rendering, both sites are occupied. (c) In this rendering, the bivalent receptor has momentarily let go from one of its sites, but still holds onto the ligand with the other. Large proteins such as antibodies continuously vibrate and “breathe,” resulting in such on-off kinetics. Bivalent or multivalent binding helps to ensure that, when one site momentarily releases the ligand, the interaction is not entirely lost, as it is in part (a). Receptor and Ligand Expression Can Vary During the Course of an Immune Response cytes and antigen-presenting cells engaged in an immune response spend significant amounts of time held together by multiple receptor-ligand interactions. Cytokine signals released by a T cell, for example, are received by a bound antigen-presenting cell at the interface between the cells, before the cytokine has time to diffuse into the tissue fluids. The secretion of cytokines into exactly the location where they are needed is facilitated by the ability of antigen receptor signaling to induce redistribution of the microtubule organizing center (MTOC) within the activated T cell. The MTOC reorientation, in turn, causes the redistribution of the secretory organelles (the Golgi body and secretory vesicles) within the T-cell cytoplasm, so that cytokines made by the T cell are secreted in the direction of the T-cell receptor, which, in turn, is binding to the antigen-presenting cell. The technical term for this phenomenon is the vectorial (directional) redistribution of the secretory apparatus. Thus, cytokines are secreted directly into the space between the activated T cell and antigen-presenting cell. Vectorial redistribution of the secretory apparatus in antigen-presenting dendritic cells also occurs upon interaction with antigen-specific T cells. This ensures that cytokines, such as IL-12, which are secreted by dendritic cells, are efficiently delivered to antigen-recognizing T cells. Figure 3-4 shows the directional secretion of the cytokine IL-12 by an antigen-presenting dendritic cell engaged in the activation of a T cell. The effector cells of the immune system, such as B cells, are also capable of presenting antigens on their cell surface for recognition by helper T cells. A B cell that has specifically recognized an antigen through its own receptor will process the antigen and express a particularly high concentration of those antigenic peptides bound to Major Histocompatibility Complex (MHC) molecules on its surface. A helper T cell, binding to an antigen presented on the surface of a B cell, One of the most striking features of the molecular logic of immune responses is that receptors for some growth factors and cytokines are expressed only on an as-needed basis. Most lymphocytes, for example, express a hetero-dimeric (two-chain), low-affinity form of the receptor for the cytokine interleukin 2 (IL-2), which initiates a signal that promotes lymphocyte proliferation. (This cytokine and its receptor are covered in more detail in Chapter 4.) This lowaffinity form of the receptor is unable to bind to IL-2 at physiological cytokine concentrations. However, when a lymphocyte is activated by antigen binding, the signal from the antigen-binding receptor causes an increase in the cell surface expression of a third chain of the IL-2 receptor (see Figure 3-3). Addition of this third chain converts the lowaffinity form of the IL-2 receptor to a high-affinity form, capable of responding to the cytokine levels found in the lymphoid organs. The functional corollary of this is that only those lymphocytes that have already bound to antigen and have been stimulated by that binding have cytokine receptors of sufficiently high affinity to respond to the physiological cytokine concentrations. By waiting until it has interacted with its specific antigen before expressing the high-affinity cytokine receptor, the lymphocyte both conserves energy and prevents the accidental initiation of an immune response to an irrelevant antigen. Local Concentrations of Cytokines and Other Ligands May Be Extremely High When considering the interactions between receptors and their ligands, it is important to consider the anatomical environment in which these interactions are occurring. Lympho- Receptors and Signaling: B and T-Cell Receptors FIGURE 3-3 The expression of some cytokine receptors is upregulated following cell stimulation. White blood cells stained with DAPI, a stain that detects DNA (blue), were either untreated (left) or stimulated using the human T-cell mitogen will therefore deliver a high concentration of cytokines directly to an antigen-specific, activated B cell. The local concentration of cytokines at the cellular interfaces can therefore be extremely high, much higher than in the tissue fluids generally. Common Strategies Used in Many Signaling Pathways A signal transduction pathway is the molecular route by which a ligand-receptor interaction is translated into a biochemical change within the affected cell. Communication of the ligand’s message to the cell is initiated by the complementarity in structure between the ligand and its receptor, which results in a binding interaction of suffi- | CHAPTER 3 69 phytohemagglutinin (right). Upon stimulation, the IL-2 receptor alpha (IL-2R␣) chain (yellow) is upregulated, increasing affinity of the IL-2 receptor for IL-2. A similar IL-2R␣ upregulation is seen upon antigen stimulation. [Courtesy R&D Systems, Inc., Minneapolis, MN, USA.] cient strength and duration to bring about a biochemical change in the receptor and/or its associated molecules. This biochemical change can, in turn, cause a diverse array of biochemical consequences in the cell (Overview Figure 3-5). The terms upstream and downstream are frequently used to describe elements of signaling pathways. The upstream components of a signaling pathway are those closest to the receptor; the downstream components are those closest to the effector molecules that determine the outcome of the pathway—for example, the transcription factors or enzymes whose activities are modified upon receipt of the signal. As disparate and complex as some of these signaling pathways may be, many share common features. In this section, in order to provide a framework for analyzing the individual pathways employed by cells of the immune system, we FIGURE 3-4 Polarized secretion of IL-12 (pink) by dendritic cells (blue) in the direction of a bound T cell (green). The higher magnification micrograph shows the secretion of packets of IL-12 through the dendritic cell membrane. [J. Pulecio et al., 2010, Journal of Experimental Medicine 207:2,719.] 70 PA R T I | Introduction 3-5 OVERVIEW FIGURE Concepts in lymphocyte signaling Coreceptor Ligand Receptor Receptor-associated molecules Receptor-associated tyrosine kinase PIP2 PIP2 PIP3 DAG PIP3 DAG Ras-GRP PLCγ PI3 kinase IP3 Ca2+ release from the endoplasmic reticulum P PKC SOS P P P Adapter P DAG P GDP/GTP exchange P P Adapter P Phosphorylation Ras Calmodulin and calcineurin activation NF-κB P P P IκB Ubiquitination and destruction MAP kinase cascade NFAT Cytoplasm NFAT Nucleus Ligand binding to receptors on a cell induces a variety of downstream effects, many of which culminate in transcription factor activation. Here we illustrate a few of the pathways that are addressed in this chapter. Binding of receptor to ligand induces clustering of receptors and signaling molecules into regions of the membrane referred to as lipid rafts (red). Receptor binding of ligand may be accompanied by binding of associated coreceptors to their own ligands, and causes the activation of receptor-associated tyrosine kinases, which phosphorylate receptor-associated proteins. Binding of downstream adapter molecules to the phosphate groups on adapter proteins creates a scaffold at the membrane that then enables activation of a variety of enzymes including phospholipase C␥ (PLC␥), PI3 kinase, and additional tyrosine kinases. PLC␥ cleaves phosphatidyl inositol bisphosphate (PIP2) to yield inositol trisphosphate (IP3), which interacts with receptors on endoplasmic reticulum vesicles to AP-1 NF-κB Gene activation cause the release of calcium ions. These in turn activate calcineurin, which dephosphorylates the transcription factor NFAT, allowing it to enter the nucleus. Diacylglycerol (DAG), remaining in the membrane after PLC␥ cleavage of PIP2, binds and activates protein kinase C (PKC), which phosphorylates and activates enzymes leading to the destruction of the inhibitor of the transcription factor NF-␬B. With the release of the inhibitor, NF-␬B enters the nucleus and activates a series of genes important to the immune system. Binding of the adapter protein Ras-GRP to the signaling complex allows for the binding and activation of the Guanine nucleotide Exchange Factor (GEF) Son of Sevenless (SOS), which in turn initiates the phosphorylation cascade of the MAP kinase pathways. This leads to the entry of a third set of transcription factors into the nucleus, and activation of the transcription factor AP-1. (Many details that are explained in the text have been omitted from this figure for clarity.) Receptors and Signaling: B and T-Cell Receptors describe some of the strategies shared by many signaling pathways. The next section provides three examples of signaling pathways used by the immune system, as well as by other organ systems. Ligand Binding Can Induce Conformational Changes in, and/or Clustering of, the Receptor The first step necessary to the activation of a signaling pathway is that the binding of the ligand to its receptors in some way induces a physical or chemical change in the receptor itself, or in molecules associated with it. In the case of many growth factor receptors, ligand binding induces a conformational change in the receptor that results in receptor dimerization (Figure 3-6). Since the cytoplasmic regions of many growth factor receptors have tyrosine kinase activity, this results in the reciprocal phosphorylation of the cytoplasmic regions of each of the receptor molecules by its dimerization partner. Other receptors undergo conformational changes upon ligand binding that result in higher orders of receptor polymerization. The receptors on B cells are membranebound forms of the antibody molecules that the B cell will eventually secrete. Two different types of antigen receptors exist on the surface of naïve B cells, which we term immunoglobulins M and D (IgM and IgD). Structural studies of the IgM form of the receptor have revealed that ligand binding induces a conformational change in a nonantigen binding part of the receptor, close to the membrane, that facilitates aggregation of the receptors into multimeric complexes and their subsequent movement into specialized membrane Growth factor Cytosol P P P Inactive receptor tyrosine kinases CHAPTER 3 71 regions. T-cell receptors similarly cluster upon antigen binding. Whether or not they undergo a conformational change upon antigen binding remains controversial. Some Receptors Require Receptor-Associated Molecules to Signal Cell Activation In contrast to growth factor receptors, which have inducible enzyme activities built into the receptor molecule itself, B- and T-cell receptors have very short cytoplasmic components and therefore need help from intracellular receptor-associated molecules to bring about signal transduction. The Ig␣/Ig␤ (CD79␣/␤) heterodimer in B cells, and the hexameric CD3 complex in T cells are closely associated with their respective antigen receptors and are responsible for transmitting the signals initiated by ligand binding into the interior of the cell (Figure 3-7). Both of these complexes have a pair of long cytoplasmic tails that contain multiple copies of the Immuno-receptor Tyrosine Activation Motif or ITAM. ITAMs are recurrent sequence motifs found on many signaling proteins within the immune system, which contain tyrosines that become phosphorylated following signal transduction through the associated receptor. Phosphorylation of ITAM-tyrosine residues then allows docking of adapter molecules, thus facilitating initiation of the signaling cascade. Other molecules associated with the B- or T-cell antigen receptors may also interact with the antigen or with other molecules on the pathogen’s surface. For example, in the case of B cells, the CD19/CD21 complex binds to complement molecules covalently attached to the antigen (see Figure 3-7a). Similarly, CD4 and CD8 molecules on T cells bind to nonpolymorphic regions of the antigen-presenting MHC molecule and aid in signal transduction (see Figure 3-7b). Finally, the co-receptor CD28 on naïve T cells must interact with its ligands CD80 (B7-1) and CD86 (B7-2) for full T-cell activation to occur. P Ligand-Induced Receptor Clustering Can Alter Receptor Location P P Ligand-induced clustering of B- and T-cell receptors slows down the rates of their diffusion within the planes of their respective cell membranes, and facilitates their movement into specialized regions of the lymphocyte membrane known as lipid rafts. These rafts are highly ordered, detergent-insoluble, cholesterol- and sphingolipid-rich membrane regions, populated by many molecules critical to receptor signaling. Moving receptors and co-receptors into the lipid rafts renders them susceptible to the action of enzymes associated with those rafts. Figure 3-8 shows how the raft-associated tyrosine kinase Lyn initiates the B-cell signaling cascade by phosphorylating the receptor-associated molecules Ig␣ and Ig␤. The tyrosine kinase Lck serves a similar role in the TCR signaling cascade. P P Tyrosine kinase domain | Kinase activity stimulated by cross-phosphorylation FIGURE 3-6 Some growth factors induce dimerization of their receptors, followed by reciprocal tyrosine phosphorylation of the receptor molecules. Many growth factor receptors possess tyrosine kinase activity in their cytoplasmic regions. Dimerization of the receptor occurs on binding of the relevant ligand and allows reciprocal phosphorylation of the dimerized receptor chains at multiple sites, thus initiating the signal cascade. As one example, stem cell factor binds to its receptor, c-kit (CD117) on the surface of bone marrow stromal cells. PA R T I 72 | Introduction C3d Antigen CD21 mIgM CD81 (TAPA-1) Class II MHC CD4 CD80 or CD86 Peptide CD19 TCR δ ε CD28 γ ε ζ ζ ss lgα/lgβ (CD79α,β) ITAMs CD3 FIGURE 3-7 Both B- and T-cell receptors require receptorassociated molecules and co-receptors for signal transduction. (a) B-cell receptors require Ig␣/Ig␤ to transmit their signal. The CD21 co-receptor, which is associated with CD19, binds to the complement molecule C3d, which binds covalently to the antigen. The interaction between CD21 on the B cell, and C3d associated with the antigen, keeps the antigen in contact with the B-cell receptor, even when the antigen-BCR binding is relatively weak. The yellow bands on the cytoplasmic regions of the receptor-associated molecules indicate Immuno-receptor Tyrosine Activation Motifs (ITAMs). Phosphorylation of tyrosine residues in these motifs allows the binding of downstream adapter molecules and facilitates signal transduction from the receptors. Both CD19 and Ig␣/Ig␤ bear intracytoplasmic ITAMs and, along with CD81 (TAPA-1), participate in downstream signaling events. (b) T-cell receptors use CD3, a complex of ␦␧, ␥␧ chains and a pair of ␨ chain molecules or a ␨␯ pair. CD4 and CD8 coreceptors bind to the non-polymorphic region of the MHC class II or class I molecules, respectively. This figure illustrates CD4 binding to MHC Class II. This binding secures the connection between the T cell and the antigen presenting cell, and also initiates a signal through CD4/8. The CD28 co-receptor provides another signal upon binding to CD80 or CD86. CD28 binding to costimulatory molecules on antigen-presenting cells is required for stimulation of naïve, but not memory, T cells. Antigen BCR Raft lgα/lgβ Lyn + Antigen Lyn – Antigen Lyn BLNK P P Syk B cells, the B-cell receptor (BCR) is excluded from the lipid rafts, which are regions of the membrane high in cholesterol and rich in glycosphingolipids. The raft is populated by tyrosine kinase signaling molecules, such as Lyn. On binding to antigen, the BCR multimerizes (clusters), and the changes in the conformation of the BCR brought BLNK Syk Signal transduction FIGURE 3-8 Lipid raft regions within membranes. In resting P P Receptor internalization about by this multimerization increase the affinity of the BCR for the raft lipids. Movement of the BCR into the raft brings it into contact with the tyrosine kinase Lyn, which phosphorylates the receptorassociated proteins Ig␣/Ig␤, thus initiating the activation cascade. A similar movement of TCRs into lipid rafts occurs upon T-cell activation. [Adapted from S. K. Pierce, 2002, Nature Reviews Immunology 2:96.] Receptors and Signaling: B and T-Cell Receptors Inactive Active SH2 SH2 P pY508 Y397 Kinase Y508 P pY397 Kinase FIGURE 3-9 Activation of Src-family kinaes. Src-family kinases are maintained in an inactive closed configuration by the binding of a phosphorylated inhibitory tyrosine residue (pY508 in this example) with an SH2 domain in the same protein. Dephosphorylation of this tyrosine opens up the molecule, allowing substrate access to the enzymatic site. Opening up the kinase also allows phosphorylation of a different internal tyrosine (pY397), which further stimulates the Src kinase activity. Tyrosine Phosphorylation is an Early Step in Many Signaling Pathways Many signaling pathways, particularly those that signal cell growth or proliferation, are initiated with a tyrosine phosphorylation event. Many of the tyrosine kinases that initiate BCR and TCR activation belong to the family of enzymes known as the Src-family kinases. Src (pronounced “sark”) family kinases have homology to a gene, c-src, that was first identified in birds. A constitutively active Rous sarcoma virus TABLE 3-1 | CHAPTER 3 73 homolog of these genes, v-src, was shown to induce a form of cancer, fibrosarcoma, in birds. The fact that a simple mutation in this viral form of a tyrosine kinase gene could result in the development of a tumor was the first hint that tyrosine kinase genes may be important in the regulation of cell proliferation and, indeed, in many aspects of cell signaling. Src-family kinases are important in the earliest stages of activation of both T and B cells: Lck and Fyn are critical to T-cell activation, and Lyn, Fyn, and Blk play the corresponding initiating roles in B cells. Since inadvertent activation of these enzymes can lead to uncontrolled proliferation—a precursor to tumor formation—it is not surprising that their activity is tightly regulated in two different but interconnected ways. Inactive Src-family tyrosine kinase enzymes exist in a closed conformation, in which a phosphorylated tyrosine is tightly bound to an internal SH2 domain (Src homology 2 domain) (Figure 3-9, Table 3-1). (SH2 domains in proteins bind to phosphorylated tyrosine residues.) For as long as the inhibitory tyrosine is phosphorylated, the Src-family kinase remains folded in on itself and inactive. In lymphocytes, the tyrosine kinase enzyme Csk is responsible for maintaining phosphorylation of the inhibitory tyrosine. However, upon cell activation, a tyrosine phosphatase removes the inhibitory phosphate and the Src-family kinase opens up into a partially active conformation. Thus, the initiating event in signal transduction is often the movement of the tyrosine kinase into a region of the cell or the membrane populated by an appropriate, activating phosphatase and distant from the inhibitory kinase Csk. Full activity is then achieved when the Src-kinase phosphorylates itself on a second, activating tyrosine residue. Selected domains of adapter proteins and their binding target motifs Domain of adapter protein Binding specificity of adapter domain Src-homology 2 (SH2) Specific phosphotyrosine (pY)–containing motifs in the context of 3–6 amino acids located carboxyterminal to the pY (An invariant arginine in the SH2 domain is required for pY binding.) Src-homology 3 (SH3) Proline-rich sequences in a left-handed polyproline helix (Proline residues are usually preceded by an aliphatic residue.) Phosphotyrosine-binding (PTB) pY-containing peptide motifs (DPXpY, where X is any amino acid) in the context of amino-terminal sequences Pleckstrin homology (PH) Specific phosphoinositides, which allow PH-containing proteins to respond to the generation of lipid second messengers generated by enzymes such as PI3 kinase WW Bind proline-rich sequences (Derive their name from two conserved tryptophan [W] residues 20–22 amino acids apart) Cysteine-rich sequences (C1) Diacylglycerol (DAG) (On association with DAG, the C1 domain exhibits increased affinity for the lipid membrane, promoting membrane recruitment of C1-containing proteins.) Tyrosine kinase-binding (TKB) Phosphotyrosine-binding domain divergent from typical SH2 and PTB domains (Consists of three structural motifs [a four-helix bundle, an EF hand, and a divergent SH2 domain], that together form an integrated phosphoprotein-recognition domain) Proline-rich Amino-acid sequence stretches rich in proline residues, able to bind modular domains including SH3 and WW domains 14-3-3- binding motifs Phosphorylated serine residues in the context of one of the two motifs RSXpSXP and RXXXpSXp [After G. A. Koretsky and P. S. Myung, 2001, Nature Reviews Immunology 1:95] 74 PA R T I | Introduction Tyrosine phosphorylation can bring about changes in a signaling pathway in more than one way. Sometimes, phosphorylation of a tyrosine residue can induce a conformational change in the phosphorylated protein itself, which can turn its enzymatic activity on or off. Alternatively, tyrosine phosphorylation of components of a receptor complex can permit other proteins to bind to it via their SH2 or phosphotyrosine binding domains (see Table 3-1), thereby altering their locations within the cell. Note that a phosphorylated tyrosine is often referred to as a “pY” residue. specificity for a particular molecular structure. In many signaling pathways, multiple adapter proteins may participate in the formation of a protein scaffold that provides a structural framework for the interaction among members of a signaling cascade. Particularly common domains in adapter proteins functioning in immune system signaling are the SH2 domain that binds to phosphorylated tyrosine (pY) residues, the SH3 domain that binds to clusters of proline residues, and the pleckstrin homology (PH) domain that binds to phosphatidyl inositol trisphosphate in the plasma membrane. Binding of adapter proteins may simply bring molecules into contact with one another, so that, for example, an enzyme may act on its substrate. Alternatively, the binding of an adapter protein may induce a conformational change, which in turn can stabilize, destabilize, or activate the binding partner. Signal transduction may induce conformational changes in proteins that in turn result in the uncovering of one or more protein domains with specific affinity for other proteins. Such interactions may be homotypic (interactions between identical domains) or heterotypic (interactions between different domains). Adapter Proteins Gather Members of Signaling Pathways It is easy to imagine the cytoplasm as an ocean of macromolecules, with little structure or organization, in which proteins bump into one another at a frequency dependent only on their overall cytoplasmic concentrations. However, nothing could be further from the truth. The cytoplasm is, in fact, an intricately organized environment, in which threedimensional arrays of proteins form and disperse in a manner directed by cellular signaling events. Many of these reversible interactions between proteins are mediated by adapter proteins as well as by interactions between members of signaling pathways with cytoskeletal components. Adapter proteins have no intrinsic enzymatic or receptor function, nor do they act as transcription factors. Their function is to bind to specific motifs or domains on proteins or lipids, linking one to the other, bringing substrates within the range of enzymes, and generally mediating the redistribution of molecules within the cell. Table 3-1 lists a number of representative adapter domains along with their binding specificities. Adapter proteins are characterized by having multiple surface domains, each of which possesses a precise binding ⫹ NH3 CH2 CH COO ⫺ Phosphorylation on Serine and Threonine Residues is also a Common Step in Signaling Pathways Whereas tyrosine phosphorylation is frequently seen at the initiation of a signaling cascade, the phosphorylation of proteins at serine and threonine residues (Figure 3-10) tends to occur at later steps in cell activation. Serine or threonine phosphorylation may serve to activate a phosphorylated enzyme, to induce the phosphorylated protein to interact with a different set of proteins, to alter the protein’s location within the cell, to protect the protein from destruction, or, in COO ⫺ COO ⫺ ⫹ H3N ⫹ CH H3N CH CH2OH CH H3C OH Tyrosine Threonine Serine Tyrosine kinase OH Serine/threonine kinases ⫹ NH3 CH2 CH COO ⫺ COO ⫺ COO ⫺ ⫹ H3N OPO32⫺ Phosphotyrosine ⫹ H3N CH CH2OPO32⫺ 3-Phosphoserine FIGURE 3-10 Phosphorylated tyrosine, serine, and threonine. CH CH H3C OPO32⫺ Phosphothreonine Receptors and Signaling: B and T-Cell Receptors some important instances, to convert the phosphorylated protein into a target for proteasomal destruction. Phosphorylation of Membrane Phospholipids Recruits PH Domain-Containing Proteins to the Cell Membrane The phospholipid Phosphatidyl Inositol bis-Phosphate (PIP2) is a component of the inner face of eukaryotic plasma membranes. However, PIP2 is much more than a structural phospholipid; it also actively participates in cell signaling. The enzyme Phosphatidyl Inositol-3-kinase (PI3 kinase), activated during the course of signaling through many immune receptors, phosphorylates PIP2 to form Phosphatidyl Inositol tris-Phosphate (PIP3) (Figure 3-11). PIP3 remains in the membrane and serves as a binding site for proteins bearing pleckstrin homology (PH) domains (see Table 3-1). Thus, this lipid phosphorylation event serves to move proteins from the cytosol to the membrane by providing them with a binding site O O O O O O⫺ OH P O O OH O O⫺ O P O⫺ O O O O O⫺ P P O⫺ O⫺ ⫺ O FIGURE 3-11 PIP2 can be phosphorylated (red) by PI3 kinase to create PIP3 during cellular activation. PIP3 then creates binding sites for proteins with PH domains at the internal face of the plasma membrane. | CHAPTER 3 75 at the inner membrane surface. Localization of proteins at the membrane brings them into contact with other members of a signaling cascade, allowing enzymes access to new substrates and enabling the modification of adapter proteins with the subsequent assembly of signaling complexes. Signal-Induced PIP2 Breakdown by PLC Causes an Increase in Cytoplasmic Calcium Ion Concentration In addition to being phosphorylated by PI3 kinase, PIP2 is also susceptible to a second type of signal-induced biochemical modification. A second enzyme (more correctly, a family of enzymes), Phospholipase C (PLC), is also activated upon antigen signaling of lymphocytes (Figure 3-12). PLC hydrolyzes PIP2, cleaving the sugar inositol trisphosphate (IP3) from the diacylglyerol (DAG) backbone. The DAG remains in the membrane, where it binds and activates a number of important signaling enzymes. IP3 is released into the cytoplasm, where it interacts with specific IP3 receptors on the surface of endoplasmic reticulum vesicles, inducing the release of stored calcium ions (Ca2⫹) into the cytoplasm. These calcium ions then bind several cellular proteins, changing their conformation and altering their activity. The calcium ion’s double positive charge and the size of its ionic radius render it particularly suitable for binding to many proteins; however, in the field of cell signaling, calmodulin (CaM) is unquestionably the most important calcium-binding protein. As shown in Figure 3-13a, CaM is a dumbbell-shaped protein, with two globular subunits separated by an ␣-helical rod. Each of the two subunits has two calcium binding sites, such that calmodulin has the capacity to bind four Ca2⫹ ions. On binding calcium, CaM undergoes a dramatic conformational change (Figure 3-13b), which allows it to bind to and activate a number of different cellular proteins. As cytoplasmic Ca2⫹ ions are used up in binding to cytoplasmic proteins, and the free cytoplasmic Ca2⫹ ion concentration drops, Ca2⫹ channel proteins begin to assemble at the membrane. Eventually the channels open to allow more calcium to flood in from the extracellular fluid and complete the activation of the calcium-regulated proteins. The assembly of these so-called store-operated calcium channel proteins is prohibited in the presence of high intracellular calcium. Many different types of ions are present in a cell, and it is reasonable to ask why eukaryotes should have evolved proteins with activities that are so responsive to intracellular Ca2⫹ ion concentrations. The answer to this question in part lies in the fact that it is relatively easy to alter the intracytoplasmic concentration of Ca2⫹ ions. The concentration of Ca2⫹ in the blood and tissue fluids is on the order of 1 mM, whereas the cytosolic concentration in a resting cell is closer to 100 nM—10,000 times lower than the extracellular concentration. This difference is maintained by an efficient (although energetically expensive) system of membrane pumps. Furthermore, higher-concentration Ca2⫹ stores exist PA R T I 76 | Introduction O O O O P HO O 2⫺ OH OPO3 OPO32⫺ OH ` O` ⫺ H O O Phosphatidylinositol 4,5-bisphosphate (PIP2) Phospholipase C O OH 2⫺ OH OPO3 ` O` H + O HO 2⫺O PO 3 OPO32⫺ OH O Diacylglycerol (DAG) (Remains associated with the membrane) Inositol 1,4,5-trisphosphate (IP3) Enters the cytoplasm and induces Ca2+ ion release from ER vesicles FIGURE 3-12 PIP2 can be hydrolyzed by PLC to create DAG and IP3. within intracellular vesicles associated with the endoplasmic reticulum and the mitochondria. Thus, any signaling event that opens up channels in the membrane of the endoplasmic reticulum, or in the plasma membrane, and allows for free flow of Ca2⫹ ions into the cytoplasm, facilitates a rapid rise in intracytoplasmic Ca2⫹ ion concentration. Ubiquitination May Inhibit or Enhance Signal Transduction The protein ubiquitin is a small, highly conserved, 76-residue, monomeric protein. Binding of the carboxy-terminal of ubiquitin to lysine residues of target proteins is often followed Ca2+ (b) (a) Ca2+ Calmodulin 1 Ca2+ +NH 2 COOH 3 Ca2+ Calciumcalmodulin complex Target protein Calmodulinbinding site 3 Calmodulin binds Ca2+ ions. Calmodulin changes conformation, becomes active. Calmodulin binds a target protein. FIGURE 3-13 The calcium-binding regulatory protein calmodulin undergoes a conformational shift on binding to calcium, which enables it to bind and activate other proteins. [Part (a) PDB ID 3CLN.] Receptors and Signaling: B and T-Cell Receptors by polymerization of a chain of ubiquitins onto selected lysine residues of the conjugated ubiquitin. This results in poly-ubiquitination of the target protein. In most instances, mono- or poly-ubiquitination serves as a mechanism by which the tagged proteins are targeted for destruction by the proteasome of the cell. However, we now know that ubiquitination on certain residues of proteins can also serve as an activating signal, and we will come across this alternative role of ubiquitin below as we learn about the activation of the NF-␬B transcription factor, as well as in Chapters 4 and 5. Frequently Encountered Signaling Pathways Analysis of the biochemical routes by which molecular signals pass from cell surface receptors to the interior of the cell has revealed that a few signal transduction pathways are used repeatedly in the cellular responses to different ligands. These pathways are utilized, in slightly different forms, in multiple cellular and organ systems, and in many species. The end result of most of these pathways is an alteration in the transcriptional program of the cell. Here we briefly describe three such pathways of particular relevance to the adaptive immune system (see Overview Figure 3-5). Each of these pathways is triggered by antigen binding to the receptor and leads to the activation of a different family of transcription factors. The generation of these active transcription factors in turn initiates the up-regulation of a cascade of genes important to the immune response, including those encoding cytokines, antibodies, survival factors, and proliferative signals. Although the conceptual framework of these signaling pathways is shared among many cell types, small modifications in the molecular intermediates and transcription factors involved allows for variations in the identities of the genes controlled by particular signals and for cell-type specific transcriptional control mechanisms. For example, the enzyme PLC exists in different forms; PLC␥1 is used in T cells and PLC␥2 in B cells. Similarly, the Nuclear Factor of Activated T cells (NFAT) family of transcription factors includes five members, some of which are expressed as differentially spliced variants, and each of which is affected by different arrays of molecular signals. Furthermore, activation of transcription from some promoters requires the binding of multiple transcription factors. For example, full expression of the gene encoding interleukin 2 (IL-2) requires the binding of AP-1, NF-␬B, and NFAT, in addition to several other transcription factors, to the interleukin promoter. In this way, the promoter can be partially activated in the presence of some of these transcription factors, but not fully active until all of them are in place. In the next section, we focus on the general principles that govern the control of these three pathways, from the receipt of an antigen signal, to the activation of a particular type of transcription factor. | CHAPTER 3 77 The PLC Pathway Induces Calcium Release and PKC Activation We have already met the essential elements of this pathway in our discussion of the enzymes that modulate PIP2 upon cell activation (see Figure 3-12), and we now know that PLC breaks down PIP2 to IP3 and DAG. But how does PLC become activated in the first place? We use T cells here as our example. As mentioned, T cells use the PLC␥1 form of the enzyme. On antigen stimulation, tyrosine phosphorylation of the ITAM residues of the CD3 receptor-associated complex results in the localization of the adapter protein LAT to the membrane (Figure 3-14). LAT in turn is phosphorylated, and PLC␥1 then binds to phosphorylated LAT. This binding ensures that PLC␥1 localizes to the cell membrane, the site of its substrate, PIP2. CD4 TCR/CD3 PIP3 PIP2 DAG Itk PLCγ1 P Lck P P LAT IP3 Endoplasmic reticulum Ca2+ Calcineurin Calmodulin (inactive) Calmodulin (active) P NFAT P Cytoplasm NFAT Nucleus Transcription factor activation FIGURE 3-14 Activation of calcineurin by binding of the calcium-calmodulin complex induces NFAT dephosphorylation and nuclear entry. 78 PA R T I | Introduction Once located at the plasma membrane, PLC␥1 is further activated by tyrosine phosphorylation mediated by the receptoractivated kinase, Lck, as well as by a second tyrosine kinase, Itk. The phosphorylated and activated PLC␥1 then mediates cleavage of PIP2 to IP3 and DAG as described above. How do these secondary mediators then function to bring about lymphocyte activation? Calcium ions, released from intracellular stores by IP3, bind to the signaling intermediate, calmodulin. Each molecule of calmodulin binds to four calcium ions, and this binding results in a profound alternation in its conformation (see Figure 3-13). The calcium-calmodulin complex then binds and activates the phosphatase calcineurin, which dephosphorylates the transcription factor NFAT (see Figure 3-14). This dephosphorylation induces a conformational change in NFAT, revealing a nuclear localization sequence, which directs NFAT to enter the nucleus and activate the transcription of a number of important T-cell target genes. Genes activated by NFAT binding include that encoding interleukin 2, a cytokine pivotally important in the control of T cell proliferation. The importance of the PLC␥1 pathway to T cell activation is illustrated by the profound immunosuppressant effects of the drug cyclosporine, which is used to treat both T-cell–mediated autoimmune disease and organ transplant rejection. In T cells, cyclosporine binds to the protein cyclophilin (immunophilin), and the cyclosporine/cyclophilin complex binds and inhibits calcineurin, effectively shutting down T-cell proliferation. PIP2 cleavage by PLC is a particularly efficient signaling strategy as both products of PIP2 breakdown are active in subsequent cellular events. DAG, the second product of PIP2 breakdown (see Figure 3-12) remains in the membrane, where it binds and activates enzymes of the protein kinase C (PKC) family. These kinases are serine/threonine kinases active in a variety of signaling pathways. Protein kinase C␪, important in T-cell signaling, only requires DAG binding, whereas its relative, protein kinase C, must also bind Ca2⫹ ions for full activation. DAG is also implicated in the Ras pathway (described in the next section). The Ras/Map Kinase Cascade Activates Transcription Through AP-1 The Ras signaling pathway was discovered initially after a mutated, viral form of the Ras protein was found to induce cancer in a rat model. Ras is a monomeric, GTP-binding protein (G protein, for short). When Ras binds to GTP, its conformation changes into an active state, in which it is capable of binding, and activating a number of serine/threonine kinases. However, the Ras protein possesses an intrinsic GTPase activity that hydrolyzes GTP to GDP, and the Ras– GDP form of the protein is incapable of transmitting a positive signal to downstream kinases (Figure 3-15). Activation of the Ras pathway is therefore dependent on the ability to maintain Ras in its GTP-bound state. Modulation between the active, GTP-bound form and the inactive, GDP-bound form is brought about by two families of enzymes. Guaninenucleotide Exchange Factors (GEFs) activate Ras by induc- Guanine nucleotide Exchange Factors (GEFs) GTP GDP GDP Inactive GTP Pi Active GTPase Activating Proteins (GAPs) FIGURE 3-15 Small monomeric G proteins, such as Ras, alter conformation depending on whether they are bound to GTP or GDP. The GDP-bound form of small G proteins such as Ras (pale green) is inactive, and the GTP-bound form (dark green) is active. Small G proteins have intrinsic GTPase activity, which is further enhanced by GTPase Activating Proteins (GAPs). Guanine nucleotide Exchange Factors (GEFs) act in an opposite manner to release GDP and promote GTP binding. ing it to release GDP and accept GTP; GTPase Activating Proteins (GAPs) inhibit the protein’s activity by stimulating Ras’s intrinsic ability to hydrolyze bound GTP. Like the PLC␥ pathway, the Ras pathway is initiated during both B and T lymphocyte activation and, again, we will use T cells as our example. DAG, released after PLC␥1-mediated PIP2 cleavage, binds and activates Ras-GRP, an adapter protein that then recruits the GEF, Son of Sevenless (SOS). SOS binds to Ras, inducing it to bind GTP, at which point Ras gains the ability to bind and activate the first member of a cascade of serine/threonine kinase enzymes that phosphorylate and activate one another. Because the members of this cascade were first identified in experiments that studied the activation of cells by mitogens (agents that induce proliferation), it is referred to as the Mitogen Activated Protein Kinase or MAP kinase cascade. The members of the cascade are held in close proximity to one another by the adapter protein KSR. In the T-cell activation form of this cascade (illustrated in Figure 3-16), RasGTP binds to the MAP kinase kinase kinase (MAPKKK) Raf. Binding of RasGTP alters the conformation of Raf and stimulates its serine/threonine kinase activity. Raf then phosphorylates and activates the next enzyme in the relay, a MAP kinase kinase (MAPKK), in this case MEK. Activated MEK then phosphorylates its substrate, Extracellular signal-Related Kinase, or Erk, a MAP kinase, which consequently gains the ability to pass through the nuclear membrane. Once inside the nucleus, Erk phosphorylates and activates a transcription factor, Elk-1, which cooperates with a second protein, serum response factor (SRF), to activate the transcription of the fos gene. The Fos protein is also phosphorylated by Erk, and along with its partner, Jun, forms the master transcription factor, AP-1. Jun is phosphorylated and activated via a slightly different form of the MAP kinase pathway. AP-1 is another of the transcription factors that facilitates the transcription of the IL-2 gene. | Receptors and Signaling: B and T-Cell Receptors CHAPTER 3 79 TCR/CD3 TCR-mediated signals DAG Pi GDP GTP Ras-GRP SOS (inactive) DAG PKCθ GTP GDP (active) Raf (MAPKKK) P MAP kinase cascade MEK (MAPKK) P TAK1 complex Bcl10 MALT1 TRAF6 ERK (MAPK) NEMO Ub P IKKα IKKβ P Cytoplasm Nucleus Elk-1 P NF-κB SRF Carma-1 Fos Ub NF-κB pathway Gene activation Fos P IκB P Cytoplasm Jun P NF-κB P Jun Fos P 5´ Gene activation Nucleus 3´ AP-1 FIGURE 3-17 Phosphorylation and ubiquitination acti- Gene activation vates IKK, which phosphorylates and inactivates I␬B, allowing translocation of NF-␬B into the nucleus. FIGURE 3-16 The Ras pathway involves a cascade of serine/threonine phosphorylations and culminates in the entry of the MAP kinase, Erk, into the nucleus where it phosphorylates the transcription factors Elk-1 and Fos. The Ras pathway is an important component in many developmental and cell activation programs. Each pathway uses slightly different combinations of downstream protein kinases, but all adhere to the same general mode of passing the signal from the cell surface through to the nucleus via a cascade of phosphorylation reactions, with the resultant activation of a new transcriptional program. PKC Activates the NF-␬B Transcription Factor NF-␬B belongs to a family of heterodimeric transcription factors, and each dimer activates its own repertoire of promoters. In resting cells, NF-␬B heterodimers are held in the cytoplasm by binding to the Inhibitor of NF-␬B (I␬B) protein. Cell activation induces the phosphorylation of these inhibitory proteins by an I␬B kinase (IKK) complex. The phosphorylated I␬B protein is then targeted for proteasomal degradation, releasing the freed NF-␬B to enter the nucleus and bind to the promoters of a whole array of immunologically important genes. NF-␬B is important in the transcription control of proteins needed for the proper functioning of many innate and adaptive immune system cell types; in general, NF-␬Bmediated transcription is associated with proinflammatory and activation events, rather than with regulatory processes. Different types of cells and receptors use various forms of protein kinase activation as well as diverse combinations of adapter molecules in the pathway leading to NF-␬B activation. However, all of these pathways culminate in the phosphorylation and subsequent destruction of the inhibitor of NF-␬B, I␬B. Below, we describe the pathway to NF-␬B activation that is triggered by TCR antigen recognition. 80 PA R T I | Introduction In T cells, activation of NF-␬B begins when DAG, generated by PLC␥1, recruits the serine/threonine kinase PKC␪ to the membrane (Figure 3-17). As mentioned above, PKC␪ is the form of the serine/threonine kinase that is used in the T-cell receptor signaling pathway. Once bound to DAG, PKC␪ is activated and phosphorylates adapter proteins including Carma1, initiating a cascade that ultimately recruits the ubiquitin ligase TRAF6. TRAF6 ubiquitinates and partially activates the IKK complex, which is made up of the regulatory protein NEMO, and two catalytic IKK subunits. The IKK complex is fully activated only when it is both phosphorylated and ubiquitinated. (Note that this is one of those instances in which ubiquitination does not result in subsequent protein degradation, but rather in its activation.) Other T cell mediated signals activate the TAK1 complex, which phosphorylates IKK, thus completing its activation and allowing it to phosphorylate I␬B. In this step, phosphorylation of I␬B signals its degradation, rather than its activation. NF-␬B is then free to move into the nucleus and activate the transcription of its target genes, including the IL2 gene. In addition to this T-cell receptor-mediated pathway of NF-␬B activation, signaling through CD28, the T-cell coreceptor, also positively controls NF-␬B activation in T cells. These three pathways together illustrate how the broad concepts introduced earlier in this chapter can be applied to understand the mechanisms underlying T-cell activation. These themes recur in numerous forms throughout the immune system. The Structure of Antibodies Having outlined some of the general concepts that underlie many different receptor-ligand interactions and signaling pathways, we now turn our attention more specifically to the antigen receptors of the adaptive immune system. On stimulation with antigen, B cells secrete antibodies with antigen-binding sites identical to those on the B-cell membrane antigen receptor. The identity between the binding sites of the secreted antibody and the membrane-bound B-cell receptor (BCR) was first demonstrated by making reagents that bound to antibodies secreted by a particular clone of B cells and showing that those reagents also bound to the receptors on the cells that had secreted the antibodies. Since working with soluble proteins is significantly easier than manipulating membrane receptor proteins, the presence of a soluble form of the receptor greatly facilitated the characterization of B-cell receptor structure. Consequently, the basic biochemistry of the B-cell receptor was established long before that of the T-cell receptor (TCR), which, unlike the BCR, is not released in a secreted form. Box 3-1: Classic Experiment details the Nobel Prize–winning experiments that established the four-chain structure of the antibody molecule. We begin this section with a discussion of antibody structure, and then go on to describe the B-cell membrane receptor and its associated signaling pathways. Secreted antibodies and their membrane-bound receptor forms belong to the immunoglobulin family of proteins. This large family of proteins, which includes both B- and T-cell receptors, adhesion molecules, some tyrosine kinases, and other immune receptors, is characterized by the presence of one or more immunoglobulin domains. Antibodies Are Made Up of Multiple Immunoglobulin Domains The immunoglobulin domain (Figure 3-18) is generated when a polypeptide chain folds into an organized series of antiparallel ␤-pleated strands. Within each domain, the ␤ strands are arranged into a pair of ␤ sheets that form a tertiary, compact domain. The number of strands per sheet varies among individual proteins. In antibody molecules, most immunoglobulin domains contain approximately 110 amino acids, and each ␤ sheet contains three to five strands. The pair of ␤ sheets within each domain are stabilized with respect to one another by an intrachain disulfide bond. Neighboring domains are connected to one another by a stretch of relatively unstructured polypeptide chain. Within the ␤ strands, hydrophobic and hydrophilic amino acids alternate and their side chains are oriented perpendicular to the plane of the sheet. The hydrophobic amino acids on one sheet are oriented toward the opposite sheet, and the two sheets within each domain are therefore stabilized by hydrophobic interactions between the two sheets as well as by the covalent disulfide bond. The immunoglobulin fold provides a perfect example of how structure determines and/or facilitates function. At the ends of each of the ␤ sheets, more loosely folded polypeptide regions link one ␤ strand to the next, and these loosely folded regions can accommodate a variety of amino acid side-chain lengths and structures without causing any disruption to the overall structure of the molecule. Hence, in the antibody molecule, the immunoglobulin fold is superbly adapted to provide a single scaffold onto which multiple different binding sites can be built, as the antigen-binding sites can simply be built into these loosely folded regions of the antigen-binding domains. These properties explain why the immunoglobulin domain has been used in so many proteins with recognition or adhesive functions. The essential domain structure provides a molecular backbone, while the loosely folded regions can be adapted to bind specifically to many adhesive or antigenic structures. Indeed, the immunoglobulin domain structure is used by many proteins besides the BCR chains. As we will see later in this chapter, the T-cell receptor is also made up of repeating units of the immunoglobulin domain. Other proteins that use immunoglobulin domains include Fc receptors; the T-cell receptor accessory proteins CD2, CD4, CD8, and CD28; the receptor-associated proteins of both the TCR and the BCR; adhesion molecules; and others. Some of these immunoglobulin domain-containing proteins are illustrated in Figure 3-19. Each of these proteins is classified as a member Receptors and Signaling: B and T-Cell Receptors (a) CL domain | CHAPTER 3 81 VL domain Loops β strands COOH NH2 Disulfide bond CDRs β-strand arrangement (b) COOH NH2 COOH NH2 CDRs FIGURE 3-18 Diagram of the immunoglobulin fold structure of the antibody light chain variable (VL) and constant (CL) region domains. (a) The two ␤ pleated sheets in each domain are held together by hydrophobic interactions and the conserved disulfide bond. The ␤ strands that compose each sheet are shown in different colors. The three loops of each variable domain show considerable variation in length and amino acid sequence; these hypervariable regions (blue) make up the antigen-binding site. of the immunoglobulin superfamily, a term that is used to denote proteins derived from a common primordial gene encoding the basic domain structure. Antibodies Share a Common Structure of Two Light Chains and Two Heavy Chains All antibodies share a common structure of four polypeptide chains (Figure 3-20), consisting of two identical light (L) chains and two identical heavy (H) chains. Each light chain is bound to its partner heavy chain by a disulfide bond between corresponding cysteine residues, as well as by noncovalent interactions between the VH and VL domains and the CH1 and CL domains. These bonds enable the formation Hypervariable regions are usually called complementarity-determining regions (CDRs). Heavy-chain domains have the same characteristic structure. (b) The ␤-pleated sheets are opened out to reveal the relationship of the individual ␤ strands and joining loops. Note that the variable domain contains two more ␤ strands than the constant domain. [Part (a) adapted from M. Schiffer et al., 1973, Biochemistry 12: 4620; reprinted with permission; part (b) adapted from A. F. Williams and A. N. Barclay, 1988, Annual Review of Immunology 6:381.] of a closely associated heterodimer (H-L). Multiple disulfide bridges link the two heavy chains together about halfway down their length, and the C-terminal parts of the two heavy chains also participate in noncovalent bonding interactions between corresponding domains. As shown in Figure 3-21, the antibody molecule forms a Y shape with two identical antigen-binding regions at the tips of the Y. Each antigen-binding region is made up of amino acids derived from both the heavy- and the lightchain amino-terminal domains. The heavy and light chains both contribute two domains to each arm of the Y, with the non–antigen-binding domain of each chain serving to extend the antigen-binding arm. The base of the Y consists of the C-terminal domains of the antibody heavy chain. CLASSIC EXPERIMENT The Elucidation of Antibody Structure FIRST SET OF EXPERIMENTS Arne Tiselius, Pers Pederson, Michael Heidelberger, and Elvin Kabat Since the late nineteenth century, it has been known that antibodies reside in the blood serum—that is, in that component of the blood which remains once cells and clotting proteins have been removed. However, the chemical nature of those antibodies remained a mystery until the experiments of Tiselius and Pederson of Sweden and Heidelberger and Kabat, in the United States, published in 1939. They made use of the fact that, when antibodies react with a multivalent protein antigen, they form a multimolecular cross-linked complex that falls out of solution. This process is known as immunoprecipitation (see Chapter 20 for modern uses of this technique). They immunized rabbits with the protein ovalbumin (the major component of egg whites), bled the rabbits to obtain an anti-ovalbumin antiserum, and then divided their antiserum into two aliquots. They subjected the first aliquot to electrophoresis, measuring the amount of protein that moved different distances from the origin in an electric field. The blue plot in Figure 1 depicts the four major protein subpopulations resolved by their technique. The first, and largest, is the albumin peak, the most abundant protein in serum, with responsibility for transporting lipids through the blood. They named the other peaks globulins. Two smaller peaks are denoted the ␣ and ␤ globulin peaks, and then a third globulin peak, ␥ globulin, clearly represents a set of proteins in high concentration in the serum. However, the most notable part of the experiment occurred when the investigators mixed their second serum aliquot with ovalbumin, the antigen. The antibodies in the serum bound to ovalbumin in a multivalent complex, which fell out of solution into a precipitate. The precipitate was then 82 removed by centrifugation. Now that they had succeeded in removing the antibodies from an antiserum, the question was, which protein peak would be affected? The black plot in Figure 1 illustrates their results. Very little protein was lost from the albumin, or the ␣ or ␤ globulin peaks. However, immunoprecipitation resulted in a dramatic decrease in the size of the ␥ globulin peak, demonstrating that the majority of their anti-ovalbumin antibodies could be classified as ␥ globulins. We now know that most antibodies of the IgG class are indeed found in the ␥ globulin class. However, antibodies of other classes are found in the ␣ and ␤ globulin peaks, which may account for the slight decrease in protein concentration found after immunoprecipitation in these other protein peaks. SECOND SET OF EXPERIMENTS Sir Rodney Porter, Gerald Edelman, and Alfred Nisonoff Knowing the class of serum protein into which antibodies fall was a start, but immunochemists next needed to figure out what antibodies looked like. The fact that they could form precipitable multivalent complexes suggested that each antibody was capable of binding to more than one site on a multivalent antigen. But the scientists still did not know how many polypeptide chains made up an antibody molecule and how many antigen-binding sites were present in each molecule. Two lines of experimentation conducted in a similar time frame on both sides of the Atlantic combined to provide the answers to these two questions. Ultracentrifugation experiments had placed the molecular weight of IgG antibody molecules at approximately 150,000 Daltons. Digestion of IgG with the enzyme papain produced three fragments, two of which were identical and a third which was clearly different (Figure 2). The third fragment, of approximately 50,000 Daltons, spontaneously formed crystals and was therefore named Fragment crystallizable, or Fc. By demonstrating that they could competitively inhibit the binding of antibodies to their antigen, the other two fragments were shown to retain the antigen-binding + − Albumin Globulins γ Absorbance The experiments that first identified antibodies as serum immunoglobulins and then characterized their familiar four-chain structure represent some of the most elegant adaptations of protein chemistry ever used to solve a biological problem. α β Migration distance FIGURE 1 Experimental demonstration that most antibodies are in the ␥-globulin fraction of serum proteins. After rabbits were immunized with ovalbumin (OVA), their antisera were pooled and electrophoresed, which separated the serum proteins according to their electric charge and mass. The blue line shows the electrophoretic pattern of untreated antiserum. The black line shows the pattern of antiserum that was first incubated with OVA to remove anti-OVA antibody and then subjected to electrophoresis. [Adapted from A. Tiselius and E. A. Kabat, 1939, Journal of Experimental Medicine 69:119, with copyright permission of the Rockefeller University Press.] BOX 3-1 capacity of the original antibody. These fragments were therefore named Fab, or Fragment antigen binding. This experiment indicated that a single antibody molecule contained two antigen binding sites and a third part of the molecule that did not participate in the binding reaction. Use of another proteolytic enzyme, pepsin, resulted in the formation of a single fragment of 100,000 Daltons, which contained two antigen-binding sites that were still held together in a bivalent molecule. Because the molecule acted as though it contained two Fab fragments, but clearly had an additional component that facilitated the combination of the two fragments into one molecule, it was named F(ab’)2. Pepsin digestion does not result in a recoverable Fc fragment, as it apparently digests it into small fragments. However, F(ab’)2 derivatives of antibodies are often used in experiments in which scientists wish to avoid artifacts resulting from binding of antibodies to Fc receptors on cell surfaces. In the second set of experiments, investigators reduced the whole IgG molecule using ␤-mercaptoethanol, in order to break disulfide bonds, and alkylated the reduced product so that the disulfide bonds could not spontaneously reform. They then used a technique called gel filtration to separate and measure the size of the protein fragments generated by this reduction and alkylation. (Nowadays, we would use SDS PAGE gels to do this experiment.) In this way, it was shown that each IgG molecule contained two heavy chains, MW 50,000 and two light chains, MW 22,000. Now the challenge was to combine the results of these experiments to create a consistent model of the antibody molecule. To do this, the scientists had to determine which of the chains was implicated in antigen binding, and which chains contributed to the crystallizable fragments. Immunologists often use immunological means to answer their questions, and this was no exception. They elected to use Fab and Fc fragments purified from rabbit IgG antibodies to immunize two separate goats. From these goats, they generated antiFab and anti-Fc antibodies, which they reacted, in separate experiments, with the heavy and light chains from the reduction and alkylation experiments. The answer was immediately clear. Anti-Fab antibodies bound to both heavy and light chains, and therefore the antigen binding site of the original rabbit IgG was made up of both heavy and light chain components. However, anti-Fc antibodies bound only to the heavy chains, but not to the light chains of the IgG molecule, demonstrating that the Fc part of the molecule was made up of heavy chains only. Finally, careful protein chemistry demonstrated that the amino-termini of the two chains resided in the Fab portion of the molecule. In this way, the familiar four-chain structure, with the binding sites at the amino-termini of the heavy and light chain pairs, was deduced from some classically elegant experiments. In 1972, Sir Rodney Porter and Gerald Edelman were awarded the Nobel Prize in Physiology and Medicine for their work in discovering the structure of immunoglobulins. Disulfide bonds SS SS SS Edelman, G. M., B. A. Cunningham, W. E. Gall, P. D. Gottlieb, U. Rutishauser, and M. J. Waxdal. 1969. The covalent structure of an entire gammaG immunoglobulin molecule. Proceedings of the National Academy of Sciences of the United States of America 63:78–85. Fleishman, J, B., R. H. Pain, and R. R. Porter. 1962 Sep. Reduction of gamma-globulins. Archives of Biochemistry and Biophysics Suppl 1:174–80. Heidelberger, M., andK. O. Pedersen. 1937. The molecular weight of antibodies. The Journal of Experimental Medicine 65:393–414. Nisonoff, A., F. C. Wissler, and L. N. Lipman. 1960. Properties of the major component of a peptic digest of rabbit antibody. Science 132:1770–1771. Porter, R. R. (1972). Lecture for the Nobel Prize for physiology or medicine 1972: Structural studies of immunoglobulins. Scandinavian Journal of Immunology 34(4):381–389. Tiselius, A., and E. A. Kabat. 1939. An electrophoretic study of immune sera and purified antibody preparations. The Journal of Experimental Medicine 69:119–131. L chain SS F(ab')2 Pepsin digestion H chain SS SS SS SS + Fc fragments Mercaptoethanol reduction Papain digestion Fab Fab + SS SS SS SS HS HS SH SH + + + SH SH SH L chains SH Fc H chains FIGURE 2 Prototype structure of IgG, showing chain structure and interchain disulfide bonds. The fragments produced by enzymatic digestion with pepsin or papain or by cleavage of the disulfide bonds with mercaptoethanol are indicated. Light (L) chains are in light blue, and heavy (H) chains are in dark blue. 83 PA R T I 84 | Introduction Immunoglobulin (IgM) V S S V S S S S S S S S S S S C S S S S C S S Ig-α/ Ig-β heterodimer V S C S S C S C S V C MHC molecules T-cell receptor C S S S S C C S S S S C V C S S S S C C CHO S S S S Class I Class II α β V S S S S S S C S S C S S S S S S C SS β2 microglobulin C S S S S C Adhesion molecules VCAM-1 S S C Poly- Ig receptor S S C T-cell accessory proteins ICAM-1 V S S V C S S C S S γ V S S V S S CD3 CD2 δ ε S S C S S S S C C V CD8 V V C S S S S C CD4 S S S S V C S S S S S S S S C S S C S S C CD64 Fcγ RI S S C S S ICAM-2 C S S C S S C S S C C S S C S S C S S C C S S LFA–3 S S C S S C S S C S S γ γ S S FIGURE 3-19 Some examples of proteins bearing immunoglobulin domains. Each immunoglobulin domain is depicted by a blue loop and, where relevant, is labeled as variable (V) or constant (C). Figure 3-21 further shows that the overall structure of the antibody molecule consists of three relatively compact regions, joined by a more flexible hinge region. The hinge region is particularly susceptible to proteolytic cleavage by the enzyme papain. Papain cleavage resolves the antibody molecule into two identical fragments that retain the antigen-binding specificity of the original antibody (shown as Fab regions in Figure 3-21), and the remaining region of the molecule, which consists of the non antigen-binding portion. This latter region, which is identical for all antibodies of a given class, crystallizes easily and was thus called the Fc region (fragment crystallizable). Each Fab region and Fc region of antibodies mediates its own particular functions during an antibody response to an antigen. The Fab regions bind to the antigen, and the Fc region of the antigen-coupled antibody binds to Fc receptors on phagocytic or cytolytic cells, or to immune effector molecules. In this way, antibodies serve as physiological bridges between an antigen present on a pathogen, and the cells or molecules that will ultimately destroy it. A family of Fc receptors exists; each Fc receptor is expressed on a different array of cells and binds to a different class of antibodies. Receptors and Signaling: B and T-Cell Receptors S H 3+ L V H1 CH2 CH2 S S O– CHO S S S S CH3 S S S CO Antigen binding Effector activity CH3 S CHO S S S L S S S S S C S C S 1 CH CL S – O Fab S S S S CO 85 Light chains NH V S S VL S 4 21 Fab 3+ Hinge S CHAPTER 3 S + 3 VH NH NH + 3 S NH | Antigenbinding site Antigenbinding site Fc 446 COO– Heavy chains COO– FIGURE 3-20 Schematic diagram of the structure of immunoglobulins derived from amino acid sequence analysis. Each heavy (dark blue) and light (light blue) chain in an immunoglobulin molecule contains an amino-terminal variable (V) region that consists of 100 to 110 amino acids and differs from one antibody to the next. The remainder of each chain in the molecule— the constant (C) regions—exhibits limited variation that defines the two light-chain subtypes and the five heavy-chain subclasses. Some heavy chains (␥, ␦, and ␣) also contain a proline-rich hinge region. The amino-terminal portions, corresponding to the V regions, bind to antigen; effector functions are mediated by the carboxy-terminal domains. The ␮ and ⑀ heavy chains, which lack a hinge region, contain an additional domain in the middle of the molecule. CHO denotes a carbohydrate group linked to the heavy chain. There are Two Major Classes of Antibody Light Chains Amino acid sequencing of antibody light chains revealed that the amino-terminal half (approximately 110 amino acids) of the light chain was extremely variable, whereas the sequence of the carboxyl-terminal half could be classified into one of two major sequence types. The N-terminal half of light chains is thus referred to as the variable, or VL, region of the light chain, and the less variable part of the sequence is termed the constant, or CL, region. The two major light chain constant region sequences are referred to as ␬ (kappa) or ␭ (lambda) chains. As more light-chain sequences were generated, it became apparent that the  chain constant region sequences could be further subdivided into four subtypes—1, 2, 3, and 4—based on amino acid substitutions at a few positions. In humans, the light chains are fairly evenly divided between the two light-chain classes; 60% of human light chains are ␬ whereas only 40% are ␭. In mice, the situation is quite different: Only 5% of mouse light chains are of the ␭ light-chain type. All light chains have a molecular weight of approximately 22 kDa. Further analysis of light-chain sequences demonstrated that, even within the variable regions of the light chain, there Hinge region = open region; site of segmental flexibility FIGURE 3-21 The three-dimensional structure of IgG. Clearly visible in this representation are the 110 amino acid immunoglobulin domains of which the molecule is constructed, along with the open hinge structure in the center of the molecule, which affords flexibility in binding to multivalent antigens. In IgG, the light chain contains two immunoglobulin domains, the heavy chain four. Fab  Antigen-binding portion of the antibody; contains paired VL/VH and CL/CH1 domains. Fc  Non–antigen-binding region of the antibody, with paired CH2/CH2 and CH3/CH3 domains. [Source: D. Sadava, H. C. Heller, G. H. Orians, W. K. Purves, and D. M. Hillis. Life: The Science of Biology, 7e (© 2004 Sinauer Associates, Inc., and W. H. Freeman & Co.), Figure 18-10 with modifications.] were regions of hypervariability. Since these hypervariable regions could be shown to interact with the bound antigen, they were renamed the complementarity-determining regions, or CDRs. Those readers with expertise in genetics will immediately identify a problem: How is it possible to encode a single protein chain with some regions that are extremely diverse and other regions that are relatively constant, while maintaining that distinction across millions of years of evolutionary drift? The solution to this puzzle involves the encoding of a single antibody variable region in multiple genetic segments that are then joined together in different combinations in each individual antibody-producing cell. This unique mechanism will be fully described in Chapter 7. There are Five Major Classes of Antibody Heavy Chains Using antibodies directed toward the constant region of immunoglobulins and amino acid sequencing of immunoglobulins derived from plasmacytoma tumor cells, investigators discovered that the sequences of the heavy-chain constant regions fall into five basic patterns. These five basic sequences have been named with Greek letters: ␮ (mu), ␦ (delta), ␥ (gamma), ⑀ (epsilon), and ␣ (alpha). Each different heavy-chain constant region is referred to as an isotype, and the isotype of the heavy 86 PA R T I TABLE 3-2 | Introduction Chain composition of the five immunoglobulin classes Heavy chain Number of CH Ig domains IgG ␥ 3 ␥1, ␥2, ␥3, ␥4 (human) ␥1, ␥2a, ␥2b, ␥3 (mouse) ␬ or ␭ None ␥2␬2 ␥2␭2 IgM ␮ 4 None ␬ or ␭ Yes (␮2␬2)n (␮2␭2)n n  1 or 5 IgA ␣ 3 ␣1, ␣2 ␬ or ␭ Yes (␣2␬2)n (␣2␭2)n n  1, 2, 3, or 4 IgE  4 None ␬ or ␭ None ⑀2␬2 ⑀2␬2 IgD ␦ 3 None ␬ or ␭ None ␦2␬2 ␦2␭2 Class* Subclasses Light chain J chain Molecular formula * See Figure 3-22 for general structures of the five antibody classes. chains of a given antibody molecule determines its class. Thus, antibodies with a heavy chain of the  isotype are of the IgM class; those with a heavy chain are IgD; those with ␥, IgG; those with ⑀, IgE; and those with ␣, IgA. The length of the constant region of the heavy chains is either 330 amino acid residues (for ␥, ␦, and ␣ chains) or 440 amino acids (for ␮ and ␧ chains). Correspondingly, the molecular weights of the heavy chains vary according to their class. IgA, IgD, and IgG heavy chains weigh approximately 55 kDa, whereas IgM and IgE antibodies are approximately 20% heavier. Minor differences in the amino acid sequences of groups of ␣ and ␥ heavy chains led to further subclassification of these heavy chains into sub-isotypes and their corresponding antibodies therefore into subclasses (Table 3-2). There are two sub-isotypes of the ␣ heavy chain, ␣1 and ␣2, and thus two IgA subclasses, IgA1 and IgA2. Similarly, there are four sub-isotypes of ␥ heavy chains, ␥1, ␥2, ␥3, and ␥4, with the corresponding formation of the four subclasses of IgG: IgG1, IgG2, IgG3, and IgG4. In mice, the four ␥ chain subisotypes are ␥1, ␥2␣, ␥2␤, and ␥3, and the corresponding subclasses of mouse IgG antibodies are IgG1, IgG2a, IgG2b, and IgG3, respectively. Further analysis revealed that the exact number, and precise positions of the disulfide bonds between the heavy chains of antibodies, vary among antibodies of different classes and subclasses (Figure 3-22). Antibodies and Antibody Fragments Can Serve as Antigens The essential principles of antibody structure were established prior to the development of the technology required to artificially generate monoclonal antibodies, and indeed much of the basic work of structure determination was completed before techniques were available for rapid DNA sequencing. As a source of homogenous antibodies, immunologists therefore turned to the protein products of antibody-secreting tumors. Plasmacytomas are tumors of plasma cells, the end-stage cell of B-cell differentiation, and the cells from that tumor are normally located in the bone marrow. When a single clone of plasma cells becomes cancerous, it is called a plasmacytoma for as long as it remains in a single bone. However, once it metastasizes into multiple bone marrow sites, the tumor is referred to as multiple myeloma. Plasmacytoma or myeloma tumors secrete large amounts of monoclonal antibodies into the serum and tissue fluids of the patients, and these antibodies can be purified in large quantities. Rather than secreting the whole antibody, some of these tumors will secrete only the light chains, or sometimes both the light chains and the whole antibodies, into the serum. The homogenous light chains secreted by these myeloma tumors are referred to as Bence-Jones proteins. In the middle to late twentieth century, tumor-derived antibodies were used to generate a great deal of information regarding antibody structure and sequence. When a means was developed by which to generate these tumors artificially in mice, data from human tumors were supplemented with information derived from laboratory-generated murine cell lines, many of which are still in use to this day. Both tumor-derived and purified serum antibodies were also used as antigens, and the anti-immunoglobulin antibodies so derived proved to be extraordinarily useful in the elucidation of antibody structure. Antibodies, or antibody fragments derived from one animal species (for example, a rabbit), can be injected into a second species | Receptors and Signaling: B and T-Cell Receptors (a) IgG (b) IgD VH VL Cγ1 CL (c) IgE VH VL CL CHAPTER 3 VH VL Cδ1 87 CL Cε1 Cε2 Hinge region Cγ2 Cε3 Cδ2 Cγ3 Cε4 Cδ3 (d) IgA (dimer) (e) IgM (pentamer) VH Cα1 VH VL VL Cµ1 CL CL Cα2 Cµ2 Disulfide bond Hinge region Cα3 J chain Cµ3 Cµ4 J chain FIGURE 3-22 General structures of the five major classes of antibodies. Light chains are shown in lighter shades, disulfide bonds are indicated by thick black lines. Note that the IgG, IgA, and IgD heavy chains contain four domains and a hinge region, whereas the IgM and IgE heavy chains contain five domains but no hinge region. The polymeric forms of IgM and IgA contain a polypeptide, called the J chain, that is linked by two disulfide bonds to the Fc region in two different monomers. Serum IgM is always a pentamer; most serum IgA exists as a monomer, although dimers, trimers, and even tetramers are sometimes present. Not shown in these figures are intrachain disulfide bonds and disulfide bonds linking light and heavy chains (see Figure 3-23). (for example, a goat), or even into another animal of the same species, in which the antibody or antibody fragment will serve as an antigen. Box 3-1 describes the series of experiments in which antibodies made against isolated proteolytic fragments of immunoglobulin molecules were used to help determine immunoglobulin structure. To understand these experiments, and many others that also use antibodies directed against whole immunoglobulins or against immunoglobulin fragments, it is important to have a good grasp of the nature of the antigenic determinants on immunoglobulin molecules. An antigenic determinant is defined as a region of an antigen that makes con- tact with the antigen-combining region on an antibody. The different types of antibodies that recognize immunoglobulin determinants are described in Table 3-3. Anti-isotype antibodies are directed against antigenic determinants present on the constant region of one particular heavy- or light-chain class of antibody, but not on any of the others. For example, an anti-isotype antibody may bind only to human  heavy chains, but not to human , ␥, ␣, or ⑀ constant regions. Alternatively, it may bind to ␬ but not to ␭ light chains. Thus, an anti-isotype antibody binds only to a single antibody constant region class or subclass. Anti-isotype antibodies are usually specific for 88 TABLE 3-3 PA R T I | Introduction Antibodies that recognize other antibodies Type of Antibody What it recognizes Anti-Fab (Fragment, antigen-binding) Papain-generated fragment of antibodies consisting of the VHCH1–VLCL domains Anti-Fc (Fragment, crystallizable) Papain-generated fragment of antibodies consisting of the paired CH2CH3 (and, for IgM and IgE, CH4) domains Anti-isotype Antigenic determinants specific to each heavy chain class Anti-allotype Antigenic determinants that are allele specific—small differences in the constant region of light and heavy chains that vary among individuals Anti-idiotype Antigenic determinants characteristic of a particular antigen combining site. Each antibody will have its own characteristic idiotypic determinants made up of residues from the heavy and light chains that contribute to the antigen-binding regions. determinants present on the constant region of antibodies from a particular species of animal, although some cross-reactivity among related species—for example, mice and rats—may occur. Some heavy- or light-chain genes occur in multiple allelic forms, and alternative allelic forms of the same isotype are referred to as allotypes. Generally, two antibodies with allotypic differences will vary in just a few residues of one of the immunoglobulin chains, and these residues constitute the allotypic determinants. Anti-allotype antibodies are generated by immunizing an individual of one species with antibodies derived from a second animal of the same species bearing an alternative form (allele) of the particular immunoglobulin gene. After an immune response has occurred, the investigator purifies or selects those antibodies that recognize the allotypic determinants from the immunized individual. Anti-allotype antibodies are used, for example, to trace particular cell populations in experiments in which B cells from a strain of animals bearing a particular allotype are transferred into a second strain that differs in immunoglobulin allotype. Finally, antibodies directed against the antigen-binding site of a particular antibody are referred to as anti-idiotypic antibodies. They can be generated by immunizing an animal with large quantities of a purified monoclonal antibody, which, by definition, bears a single antigen-binding site. Those antibodies from the immunized animal that recognize the antigen-binding site of the original antibody can then be purified. Anti-idiotypic antibodies were used in the initial experiments that proved that the B-cell receptor had the same antigen-binding site as the antibody secreted by that B cell, and they are now used to follow the fate of B cells bearing a single receptor specificity in immuno-localization experiments. As described above, treatment of whole antibody molecules with the enzyme papain cleaves the antibody molecule at the hinge region and releases two antigen-binding Fab fragments and one Fc fragment per antibody molecule. Anti-Fab antibodies and anti-Fc antibodies are made by immunizing a different species of animal from that which provided the antibody fragments with Fab or Fc fragments, respectively. Each of the Domains of the Antibody Heavy and Light Chains Mediate Specific Functions Antibodies protect the host against infection, by binding to pathogens and facilitating their elimination. Antibodies of different heavy-chain classes are specialized to mediate particular protective functions, such as complement activation (Chapter 6), pathogen agglutination, or phagocytosis (Chapter 13), and each different domain of the antibody molecule plays its own part in these host defense mechanisms. Here we briefly examine the structure of each of the antibody heavychain domains in turn, beginning with the CH1 and CL domains, which are those proximal to the VH and VL domains respectively (see Figure 3-20). CH1 and CL Domains As discussed above, the strength (avidity) of receptor binding to antigen is greatly enhanced by receptor multivalency. Antibodies have evolved to take advantage of this property by employing two antigen-binding sites, each of which can bind to individual determinants on multivalent antigens, such as are found on bacterial surfaces. The CH1 and CL domains serve to extend the antigen-binding arms of the antibody molecule, facilitating interactions with multivalent antigens and maximizing the ability of the antibody to bind to more than one site on a multivalent antigen. An interchain disulfide bond between these two domains holds them together and may facilitate the ability of some VH-VL pairs to hold on to one another and form a stable combining site. Receptors and Signaling: B and T-Cell Receptors IgG1 IgG2 IgG3 | CHAPTER 3 89 IgG4 Disulfide bond FIGURE 3-23 General structure of the four subclasses of human IgG, which differ in the number and arrangement of the interchain disulfide bonds (thick black lines) linking the heavy chains. A notable feature of human IgG3 is its 11 interchain disulfide bonds. The Hinge Regions The ␥, ␦, and ␣ heavy chains contain an extended peptide sequence between the CH1 and CH2 domains that has no homology with the other domains (see Figures 3-20 and 3-22). This so-called hinge region is rich in proline residues, rendering it particularly flexible, and as a consequence, the two antigen-binding arms of IgG, IgD, and IgA antibodies can assume a wide variety of angles with respect to one another, which facilitates efficient antigen binding. The extended nature of the amino acid chain in the hinge region contributes to the vulnerability of this part of the molecule to protease cleavage, a vulnerability that was ingeniously exploited in the early experiments that characterized antibody structure (see Box 3-1). In addition to these proline residues, the hinge region also contains a number of cysteines, which participate in heavy-chain dimerization. The actual number of interchain disulfide bonds in the hinge region varies considerably among different heavy-chain classes and subclasses of antibodies (Figure 3-23) as well as between species. Lacking a hinge region, the heavy chains of IgE make their inter-heavy chain disulfide bonds between the CH1 and CH3 domains. In IgM, disulfide bonds bridge the pairs of heavy chains at the level of CH3 and CH4. Although ␮ and ␧ chains have no hinge regions, they do have an additional immunoglobulin domain that retains some hingelike qualities. Carbohydrate Chains The two CH2 domains of ␣, ␦, and ␥ chains and the two CH3 domains of ␮ and ␧ chains are separated from their partner heavy-chain domains by oligosaccharide side chains that prevent the two heavy chains from nestling close to one another (Figure 3-24). As a result, the paired domains are significantly more accessible to the aqueous environment than other constant region domains. This accessibility is thought to contribute to the ability of IgM and IgG antibodies to bind to complement components. Immunoglobulins are in general quite extensively glycosylated, and some antibodies even have carbohydrates attached to their light chains. FIGURE 3-24 Carbohydrate residues, shown in pink, prevent close contact between the CH2 domains of human IgG1. [PDB ID 1IGT.] The Carboxy-Terminal Domains The five classes of antibodies can be expressed as either membrane or secreted immunoglobulin. Secreted antibodies 90 PA R T I | Introduction Primary transcript AAA Alternative splicing Processed transcripts AAA Secreted form Hydrophilic segment AAA Membrane-bound form lgα/lgβ Membrane-bound segments FIGURE 3-25 Membrane vs. secreted forms of immunoglobulin are created by alternative mRNA splicing. The dark blue boxes correspond to the mRNA transcript sections of the rearranged variable region sequences, and the light-blue boxes correspond to the parts of the mRNA transcript corresponding to the individual constant region domains of the heavy-chain genes. The IgM gene, shown here, contains one variable region and four constant region exons. The pink segment represents the transcript encoding the C-terminal portion of the secreted immunoglobulin. The green segments represent the transcript encoding the C-terminal portions of the membrane-bound immunoglobulin receptor, including the transmembrane and cytoplasmic regions. Alternative splicing creates the two different types of immunoglobulin molecules. Membrane-bound and secreted immunoglobulins on any one B cell share the identical antigen-binding regions and most of the heavy-chain sequences. have a hydrophilic amino acid sequence of various lengths at the carboxyl terminus of the final CH domain. In membranebound immunoglobulin receptors, this hydrophilic region is replaced by three regions (Figure 3-25): X-ray Crystallography Has Been Used to Define the Structural Basis of AntigenAntibody Binding • An extracellular, hydrophilic “spacer” sequence of approximately 26 amino acids • A hydrophobic transmembrane segment of about 25 amino acids • A very short, approximately three amino acid, cytoplasmic tail Crystallography has been used to explore the nature of antigen-antibody binding for a large number of antibodies and has demonstrated that either or both chains might provide the majority of contact residues with any one antigen. Thus, some antibodies bind to the antigen mainly via contacts with heavy-chain variable region residues, with the light chain merely providing structural support to the binding site; for other antigen-antibody interactions, the opposite is true. Still other antibodies use residues from both chains to directly contact the antigen. In most cases, though, contacts between antigen and antibody occur over a broad face and protrusions or depressions in the antibody surface are likely to be matched by complementary depressions or protrusions in the antigen. Figure 3-26 illustrates the binding of an antibody to the tip of the influenza hemagglutinin molecule. In another well-studied case of an antibody binding to the protein lysozyme, the surface area of interaction was shown to be quite large, ranging in different antibodylysozyme pairs from 650 to 900 square Angstroms. Given the tight binding between an antibody and its complementary antigen, it should not be surprising that, at least in some cases, binding of antigen to antibody induces a conformational change in the antibody, which can be visualized by x-ray crystallography (Figure 3-27). B cells express different classes of membrane immunoglobulin at particular developmental stages and under different stimulatory conditions. Immature, pre-B cells express only membrane IgM. Membrane IgD co-expression along with IgM is one of the markers of differentiation to a fully mature B cell that has yet to encounter antigen. Following antigen stimulation, IgD is lost from the cell surface, and the constant region of the membrane and secreted immunoglobulin can switch to any one of the other isotypes. The antibody class secreted by antigen-stimulated B cells is determined by cytokines released by T cells and antigen presenting cells in the vicinity of the activated B cell. Antibodies of different heavy-chain classes have selective affinities for particular cell surface Fc receptors, as well as for components of the complement system. The effector functions of particular antibody classes are further discussed in Chapters 6 and 13. Receptors and Signaling: B and T-Cell Receptors (a) | CHAPTER 3 91 (b) Antigen Antibody FIGURE 3-26 Computer simulation of an interaction between antibody and influenza virus antigen. (a) The antigen (yellow) is shown interacting with the antibody molecule; the variable region of the heavy chain is red, and the variable region of the light chain is blue. (b) The complementarity of the two molecules is revealed by separating the antigen from the antibody by 8 Å. [Based on x-ray crystallography data collected by P. M. Colman and W. R. Tulip. From G.J.V.H. Nossal, 1993. Scientific American 269(3):22.] H3 L1 H2 H1 L2 L3 FIGURE 3-27 Conformational change can occur on binding of antigen to antibody. This figure shows a complex between a peptide derived from HIV protease and an Fab fragment from an anti-protease antibody. The peptide is shown in black. The red line shows the structure of the Fab fragment before it binds the peptide, and the blue line shows its structure after binding. There are significant conformational changes in the CDRs of the Fab on binding to the antigen. These are especially pronounced in the light chain CDR1 (L1) and the heavy chain CDR3 (H3). [From J. Lescar et al., 1997, Journal of Molecular Biology 267:1207; courtesy of G. Bentley, Institute Pasteur.] Signal Transduction in B Cells Having described the structure of antibody molecules and their membrane-bound form, the B-cell receptor (BCR), we now turn our attention to BCR function. Recall that the structure of the BCR is identical to that of the antibodies that the cell will secrete on antigen stimulation, with the exception of the C-terminal portion of the heavy chain; this portion is modified so as to anchor the receptor into the B-cell plasma membrane. Prior to antigen recognition, mature B cells residing in the secondary lymphoid tissues, such as the spleen or lymph nodes, express membrane-bound forms of both IgM and IgD. As described above, the cytoplasmic tail of the BCR heavy chain is extremely short—only three amino acids—and so one of the puzzles that had to be solved by those exploring BCRmediated antigen activation was how such a short cytoplasmic tail could efficiently pass a signal into the cytoplasm. This problem was solved when co-immunoprecipitation experiments revealed that each BCR molecule was noncovalently associated with a heterodimer, Ig␣/Ig␤ (see Figure 3-7a), that is responsible for transducing the antigen signal into the interior of the cell. Recall that Ig␣/Ig␤ chains contain ITAMs, which include tyrosine residues that become phosphorylated on activation through the receptor, and serve as docking residues for downstream signaling components. The BCR is therefore structurally and functionally divided into two components: a recognition component (the immunoglobulin receptor) and a signal transduction component (Ig␣/Ig␤). In previous sections of this chapter, we described some general principles of signal transduction and outlined three commonly employed signal transduction pathways (see Overview Figure 3-5). In this section, we show how those principles apply to the intracellular events that follow upon antigenreceptor interactions in B cells. Antigen Binding Results in Docking of Adapter Molecules and Enzymes into the BCR-Ig␣/Ig␤ Membrane Complex Antigen binding induces conformational alterations in the BCR, which expose regions in the C␮4 domains of the 92 PA R T I | Introduction CD21 CD19 PIP3 Lyn PIP3 P P P P P P P P P Syk P Akt PDK1 BCAP PI3K lgα/lgβ P PIP3 P Lyn PLCγ2 Btk P P BLNK VAV P P Grb-2 Phosphorylation and inactivation Rac/Rho/ cdc42 SOS Ca2+ P PI3K Cytoskeletal reorganization AntiBax Bad apoptotic Ras PKC Calmodulin Cell survival MAP kinase cascade Calcineurin Cytoplasm NFAT NF-κB AP-1 Nucleus Gene activation FIGURE 3-28 Signal transduction pathways emanating from the BCR. Antigen-mediated receptor clustering into the lipid raft regions of the membrane leads to src-family kinase phosphorylation of the co-receptors Ig␣/Ig␤ and CD19, the adapter proteins BLNK and BCAP, and the tyrosine kinase, Syk. BCAP and CD19 recruit PI3 kinase to the membrane with generation of PIP3 and subsequent localization of PDK1 and Akt to the membrane. Phosphorylation by Akt enhances cell survival and leads to activation of the transcription factors NF-␬B and CREB as described. Activation of the B cell isoform of PLC, PLC␥2 occurs on binding to the membrane-localized adapter, BLNK and phosphorylation by Syk, resulting in the generation of DAG and IP3, with activation of the NFAT and NF-␬B pathways as described in the text. MAP kinase pathways are activated through the binding of Grb2 to BLNK, with subsequent activation of Ras and through the activation of the GEF protein VAV which activates Rac. receptor heavy chains. Interactions between neighboring receptor molecules through these domains seed receptor oligomerization (formation of small clusters of antigenreceptor complexes) and subsequent movement of these receptor clusters into specialized lipid raft regions of the B-cell membrane. There, the ITAM residues of Ig␣/Ig␤ are brought into close proximity with the Src-family kinases Lyn (Figure 3-28), Fyn, and Blk. Tyrosine phosphorylation of the Ig␣/Ig␤ ITAM residues by these Src-family kinases, particularly Lyn, then provides attachment sites for the adapter protein BLNK and an additional tyrosine kinase, Syk, which is phosphorylated and activated by the Src-family kinases. Syk then phosphorylates BLNK, providing docking sites for multiple downstream components of the signaling pathway. The adapter protein BCAP and CD19, the B-cell co-receptor, are also phosphorylated by these tyrosine kinases and serve to recruit the enzyme PI3 kinase to the plasma membrane. [Adapted from M. E. Conley et al. 2009. Annual Reviews of Immunology 27:199–227.] B Cells Use Many of the Downstream Signaling Pathways Described Above The downstream signaling pathways used by the BCR will now be familiar. The tyrosine kinases Syk and Btk together phosphorylate and activate PLC␥2, which hydrolyzes PIP2, as Receptors and Signaling: B and T-Cell Receptors described above. The resultant increase in intracytoplasmic Ca2⫹ concentrations induces the activation of calcineurin and the movement of NFAT into the nucleus (see Figure 3-14). The other product of PIP2 hydrolysis, DAG (see Figure 3-12), remains in the membrane and binds the B-cell isoform of protein kinase C, leading to the phosphorylation and release of the NF-␬B inhibitor as described above for T cells (see Figure 3-17). This results in the nuclear localization and activation of NF␬B. Additional downstream effector pathways elicit the many other changes that take place upon B-cell activation. For example, PI3 kinase, now localized at the membrane, phosphorylates PIP2 to PIP3 (see Figure 3-11), allowing the recruitment of the PH domain–containing proteins PDK1 and Akt. On phosphorylation by the serine-threonine kinase PDK1, Akt promotes cell survival by phosphorylating and inactivating pro-apoptotic molecules such as Bax and Bad. It also phosphorylates and further activates the transcription factors NF-␬B and CREB, which both support proliferation, differentiation, and survival functions of the activated B cells. The MAP kinase pathway (see Figure 3-16) is also activated during B-cell activation. Grb2 attachment to BLNK | CHAPTER 3 93 brings about the binding of the GEF SOS, followed by the binding and activation of Ras, as described above. Similarly, Rac, another small monomeric G protein of the Ras family, is activated by binding to the GEF protein Vav. Ras and Rac are both small G proteins that act to initiate MAP kinase signaling pathways. As described above, activation of the MAP kinase pathway culminates in the expression of the phosphorylated transcription factor Elk. In B cells, Elk promotes the synthesis of the transcription factor Egr-1, which acts to induce alterations in the cell-surface expression of important adhesion molecules, and ultimately serves to aid B lymphocyte migration into and within the secondary lymphoid tissues. Downstream effectors from Rac promote actin polymerization, further facilitating B-cell motility. In conclusion, antigen binding at the BCR leads to multiple changes in transcriptional activity, as well as in the localization and motility of B cells, which together result in their enhanced survival, proliferation, differentiation, and eventual antibody secretion. As exemplified in Box 3-2, defects in proteins involved in B-cell signaling can lead to immunodeficiencies. Box 3-2 CLINICAL FOCUS Defects in the B-Cell Signaling Protein Btk Lead to X-Linked Agammaglobulinemia The characterization of the proteins necessary for B- and T-cell signaling opened up new avenues of exploration for clinicians working with patients suffering from immunodeficiency disorders. Clinicians and immunogeneticists now work closely together to diagnose and treat patients with immunodeficiencies, to the benefit of both the clinical and the basic sciences. Characterization of the genes responsible for disorders of the immune system is complicated by the fact that antibody deficiencies may result from defective genes encoding either T- or B-cell proteins (since T cells provide helper factors necessary for B-cell antibody production), or even from mutations in genes encoding proteins in stromal cells important for healthy B-cell development in the bone marrow. However, no matter what the cause, all antibody deficiencies manifest clinically in increased susceptibility to bacterial infections, particularly those of the lung, intestines, and (in younger children) the ear. In 1952, a pediatrician, Ogden Bruton, reported in the journal Pediatrics the case of an eight-year-old boy who suffered from multiple episodes of pneumonia. When the serum of the boy was subjected to electrophoresis, it was shown to be completely lacking in serum globulins, and his disease was therefore named agammaglobulinemia. This was the first immunodeficiency disease for which a laboratory finding explained the clinical symptoms, and the treatment that Bruton applied— administering subcutaneous injections of gamma globulin—is still used today. As similar cases were subsequently reported, it was noted that most of the pediatric cases of agammaglobulinemia occurred in boys, whereas when the disease was reported in adults, both men and women appeared to be similarly affected. Careful mapping of the disease susceptibility to the X chromosome resulted in the pediatric form of the disease being named XLA, for X-linked agammaglobulinemia. With the characterization of the BCR signal transduction pathway components in the 1980s and 1990s came the opportunity to define which proteins are damaged or lacking in particular immunodeficiency syndromes. In 1993, 41 years after the initial description of the disease, two groups independently reported that many cases of XLA resulted from mutations in a cytoplasmic tyrosine kinase called Bruton’s tyrosine kinase, or Btk; at this point, we now know that fully 85% of patients affected with XLA have mutations in the Btk gene. Btk is a member of the Tec family of cytoplasmic tyrosine kinases, which are predominantly expressed in hematopoietic cells. Tec family kinases share a C-terminal kinase domain, preceded by SH2 and SH3 domains, a proline-rich domain, and an amino-terminal PH domain, capable of binding to PIP3 phospholipids generated by PI3 kinase activity. Btk is expressed in both B cells and platelets and is activated following signaling through the BCR, the pre-BCR (which is expressed in developing B cells), the IL-5 and IL-6 receptors, and also the CXCR4 chemokine receptor. Its involvement in pre-BCR signaling explains why (continued) 94 PA R T I CLINICAL FOCUS | Introduction Box 3-2 (continued) children with XLA suffer from defective B-cell development. Following activation, Btk moves to the inner side of the plasma membrane, where it is phosphorylated and partially activated (see Figure 3-28). Activation is completed when it autophosphorylates itself at a second phosphorylation site. Btk binds to the adapter protein BLNK, along with PLC␥2. Btk then phosphorylates and activates PLC␥2, leading, as described, to calcium flux and activation of the NF-␬B and NFAT pathways. Btk therefore occupies a central position in B-cell activation, and it is no surprise that mutations in its gene result in such devastating consequences. Over 600 different mutations have been identified in the btk gene, with the vast majority of these resulting from single base pair substitutions, or the insertion or deletion of less than five base pairs. As for other X-linked mutations that are lethal without medical intervention, XLA disease is maintained in the population by the generation of new mutations. Patients with XLA are usually healthy in the neonatal (immediately after birth) period, when they still benefit from maternal antibodies. However, recurrent bacterial infections begin between the ages of 3 months and 18 months, and currently the mean age at diagnosis in North America is 3 years. XLA is a so-called “leaky” defect; almost all children with mutations in btk have some serum immunoglobulin, and a few B cells in the peripheral circulation. The prognosis for patients who are treated with regular doses of gamma globulin has improved dramatically over the last 25 years, with the use of prophylactic injections of gamma globulin. The B cells in patients with XLA have a distinctive phenotype that can be used for diagnostic purposes. CD19 expression is low and variable in patients with XLA, whereas membrane IgM expression, normally variable in mature B cells, is relatively high and consistent in patients with XLA. This phenotype can be observed in Figure 1, which shows the flow cytometric profiles of a normal control individual (left two graphs) and a patient with a defective btk gene (right two graphs). In the top two plots, we note that the XLA patient has very few CD19⫹ B cells compared with the control and that the levels of CD19 on the surface of those B cells that do exist are lower than those of the control cells. In the lower two plots, we note that, although there are fewer CD19⫹ B cells overall in the btk-compromised patient, all of those CD19⫹ B cells have relatively high levels of surface IgM, whereas the levels of membrane IgM are much more variable in the healthy controls. Bruton, O. C. 1952. Agammaglobulinemia. Pediatrics 9:722–728. Conley, M. E., et al. 2009. Primary B cell immunodeficiencies: Comparisons and contrasts. Annual Review of Immunology 27:199–227. Tsukuda, S., et al. 1993. Deficient expression of a B cell cytoplasmic tyrosine kinase in human X-linked agammaglobulinemia. Cell 72:279–290 Vetrie, D., et al. 1993. The gene involved in X-linked agammaglobulinemia is a member of the src family of protein tyrosine kinases. Nature 361:226–233. Control CD19 Btk CD19 Control antibody CD19 Control antibody CD19 sIgM sIgM FIGURE 1 FACS profiles of a normal individual and an XLA patient. [Adapted from Conley et al. 2009. Annual Review of Immunology 27:199.] B Cells Also Receive Signals Through Co-Receptors The immunoglobulin receptor on the B cell membrane is noncovalently associated with three transmembrane molecules: CD19, CD21, and CD81 (TAPA-1) (see Figure 3-7a). Antigens are sometimes presented to the BCR already covalently bound to complement proteins, in particular to the complement component C3d. (The complement cascade is discussed in Chapter 6.) The B-cell co-receptor CD21 specifically binds to C3d, on C3d-coated antigens. This coengagement of the BCR and CD21 brings the co-receptor and the BCR into close apposition with one another. When this happens, tyrosine residues on the cytoplasmic face of the co-receptor become phosphorylated by the same enzymes that phosphorylate the ITAMs on Ig␣/Ig␤, providing sites of attachment for PI3 kinase. As illustrated in Figure 3-28, localization of PI3 kinase to the co-receptor enhances Receptors and Signaling: B and T-Cell Receptors both cell survival and the alterations in the transcriptional program that accompany cell activation. CD19 also serves as an additional site of recruitment of PLC␥. T-Cell Receptors and Signaling T cells bind complex antigens made up of peptides located in the groove of membrane-bound MHC proteins. When the T-cell receptor makes contact with its MHC-peptide antigen on the surface of an antigen-presenting cell, the two cell membranes are brought into close apposition with one another. This adds an additional layer of complexity to the process of T-cell activation that finds no parallel in B-cell signaling. Notwithstanding this additional complexity, the events of T-cell activation still unfold according to a mix of the strategies described above, and bear many similarities to B-cell receptor signaling. Here, we briefly describe T-cell receptor structure and then turn to a characterization of the signaling routes through this receptor. In Box 3-3, we provide a description of the experiments that resulted in the isolation and characterization of the ␣␤TCR. The T-Cell Receptor is a Heterodimer with Variable and Constant Regions There are two types of T-cell receptors, both of which are heterodimers (dimers made up of two different polypeptides). The majority of recirculating T cells bear ␣␤ heterodimers, which bind to ligands made up of an antigenic peptide presented in a molecular groove on the surface of a type I or type II MHC molecule. A second subset of T cells instead expresses a heterodimeric T-cell receptor composed of a different pair of protein chains, termed ␥ and ␦. T cells bearing ␥␦ receptors have particular localization patterns (often in mucosal tissues) and some ␥␦ T cells recognize different types of antigens from those bound by ␣␤ T cells. Although some ␥␦ T cells recognize conventional MHC-presented peptide antigens, other ␥␦ T cells bind lipid or glycolipid moieties presented by noncanonical MHC molecules. Yet other ␥␦ T-cell clones appear to recognize self-generated heat shock proteins or phosphoantigens derived from microbes. A unified theory of the precise nature of antigens recognized by ␥␦ T cells remains elusive. However, this ability of ␥␦ T cells to break the rules of MHC restriction may account for the evolution of a slight difference in the angle between the antigen-binding and constant regions of the T-cell receptor, which is apparent in an x-ray crystallographic analysis of the two types of receptor (Figure 3-29). Notwithstanding these functional differences in the ␣␤ versus the ␥␦ receptors, their essential biochemistry is quite similar. For the remainder of this chapter, we focus on the structure and signaling of the dominant ␣␤ TCR type, recognizing | CHAPTER 3 95 that minor differences may exist between the two types of receptors and their downstream components. Although the TCR is not an immunoglobulin per se, the TCR proteins are members of the immunoglobulin superfamily of proteins and therefore the domain structures of ␣␤ and ␥␦ TCR heterodimers are strikingly similar to those of the immunoglobulins (see Figure 3-19). The ␣ chain has a molecular weight of 40–50 kDa, and the ␤ chain’s is 40–45 kDa. Like the antibody light chains, the TCR chains have two immunoglobulin-like domains, each of which contains an intrachain disulfide bond spanning 60 to 75 amino acids. The C␣ domain of the TCR differs from most immunoglobulin domains in that it possesses only a single ␤ sheet, rather than a pair, and the remainder of the sequence is more variably folded. The amino-terminal (variable) domain in both chains exhibits marked sequence variation, but the sequences of the remainder of each chain are conserved (constant). Each of the TCR variable domains has three hypervariable regions, which appear to be equivalent to the complementarity-determining regions (CDRs) in immunoglobulin light and heavy chains. A fourth hypervariable region on the TCR␤ chain does not appear to contact antigen, and its functional significance is therefore uncertain. At the C-terminal end of the constant domain, each TCR chain contains a short connecting sequence, in which a cysteine residue forms a disulfide link with the other chain of the heterodimer. C-terminal to this disulfide is a transmembrane region of 21 or 22 amino acids, which anchors each chain in the plasma membrane. The transmembrane domains of the TCR ␣ and ␤ chains are unusual in that they each contain positively charged amino acid residues that promote interaction with corresponding negatively charged residues on the chains of the signaltransducing CD3 complex. Finally, like BCRs, each TCR chain contains only a very short cytoplasmic tail at the carboxyl-terminal end. V domains C domains ␥␦ TCR 111⬚ ␣␤ TCR 147⬚ FIGURE 3-29 Comparison of the crystal structures of ␥␦ and ␣␤ TCRs. The difference (highlighted with black lines) in the elbow angle between the ␥␦ and ␣␤ forms of the TCR. CLASSIC EXPERIMENT The Discovery of the ␣␤ T-Cell Receptor Once scientists had established that the BCR was simply a membranebound form of the secreted antibody, the elucidation of BCR structure became a significantly more tractable problem. 1 However, investigators engaged in characterizing the T-cell receptor (TCR) did not enjoy the same advantage, as the TCR is not secreted in soluble form. Understanding of TCR biochemistry Generation of a T cell hybridoma with known antigen specificity 4 Production of antibodies that bind to TCR on the T cell hybridoma Immunize a new mouse with cells from an OVA-specific T cell hybridoma Immunize mouse with ovalbumin (OVA) Wait several days Remove lymph nodes Culture lymph node cell T cells with OVA Wait several days Isolate spleen 5 2 therefore lagged behind that of the BCR until the 1980s, when an important scientific breakthrough—the ability to make monoclonal antibodies from artificially constructed B-cell tumors, or Fuse B cells from spleen with long-term B cell line Add polyethylene glycol to induce fusion of antigen-specific T cells with long-lived T cell line 6 3 Dilute out fused cells so that each well contains a single T cell hybridoma. Allow cells to divide to form clones and test each clone for its ability to secrete IL-2 when stimulated with ovalbumin peptides. Grow up individual clones of T cell hybridomas that recognize OVA. Selection and expansion of those long-term B cell clones that secrete monoclonal antibodies that bind to the T cell hybridoma Collect monoclonal antibodies that bind to T cell lines FIGURE 1 The generation of antibodies specific for the TCR 96 BOX 3-3 hybridomas—made the analysis of the TCR more technically feasible. A hybridoma is a fusion product of two cells. B-cell hybridomas are generated by artificially fusing antibody-producing, short-lived lymphocytes with long-lived tumor cells in order to generate long-lived daughter cells secreting large amounts of monoclonal antibodies. (The term monoclonal refers to the fact that all of the cells in a given hybridoma culture are derived from the single clone of cells, and therefore carry the same DNA; details of the technology are described in Chapter 20.) Although this technique was first developed for the generation of long-lived B cells, scientists working in the laboratory of John Kappler and Philippa Marrack also applied it to T lymphocytes. The researchers began by immunizing a mouse with the protein ovalbumin (OVA), allowing OVA-specific T cells to divide and differentiate for a few days, and then harvesting the lymph nodes from the immunized animal. To enrich their starting population with as many OVAspecific T cells as possible, they cultured the harvested lymph node cells in vitro with OVA for several hours (Figure 1, step 1). 7 After some time in culture, they fused these activated, OVA-specific T cells with cells derived from a T-cell tumor (Figure 1, step 2), thus generating a number of long-lived T-cell hybridoma cultures that recognized OVA peptides in the context of the MHC of the original mouse, an MHC allele called H-2d. They then diluted out the fused cells in each culture, generating several T-cell hybridoma lines in which all the cells in an individual hybridoma line derived from the product of a single fusion event (Figure 1, step 3). This is referred to as cloning by limiting dilution. In this way, they isolated a T-cell hybridoma that expressed a TCR capable of recognizing a peptide from OVA, in the context of MHC Class 2 proteins from mice of the H-2d strain. These T cells could now be used as antigens and injected into a mouse (Figure 1, step 4). The spleen of this mouse was removed a few days later, and the mouse B cells were fused with B lineage tumor cells (Figure 1, step 5). After culturing to stabilize the hybrids, the investigators cloned the B-cell hybridomas and selected a B-cell hybridoma line that produced monoclonal antibodies that bound specifically to the T-cell hybridoma (Figure 1, step 6). Most important, these antibodies interfered with the T cell’s ability to recognize its cognate antigen (Figure 1, step 7). The fact that this monoclonal antibody inhibited TCR antigen binding suggested that the antibody was binding directly to the receptor, and competing with the antigen for TCR binding. They then used these antibodies to immunoprecipitate the TCR from detergent-solubilized membrane preparations and purify the TCR protein (Figure 1, step 8). Concurrent with these experiments, the laboratories of Stephen Hedrick and Mark Davis at NIH and Tak Mak in Toronto had been making headway searching for the genes encoding the T-cell receptor. These experiments, as well as subsequent work by Susumu Tonegawa, which completed the identification of the TCR genes, are described in detail in Chapter 7. Haskins et al. 1983. The major histocompatibility complex-restricted antigen receptor on T cells. I. Isolation with a monoclonal antibody. The Journal of Experimental Medicine 157:1149–1169. Haskins et al. 1984. The major histocompatibility complex-restricted antigen receptor on T cells. Annual Review of Immunology 2:51–66. Identification of monoclonal antibody that interferes with antigen recognition by T cell hybridoma cells Mixing T cell hybridoma with OVA-presenting cells in absence of antibody results in T hybridoma proliferation OVA Mixing T cell hybridoma with OVA-presenting cells in presence of antibody results in no T hybridoma proliferation OVA 8 The two chains of the αβ receptor can be seen in lane 1 of the gel, running on top of one another at approximately 40-45 kDa. The two higher molecular weight bands seen in this lane were shown to represent non-specific bands. 45 kDa αβ TCR bands 97 PA R T I 98 | Introduction The T-Cell Signal Transduction Complex Includes CD3 Just as B-cell signaling requires the participation of the Ig␣/Ig␤ signal transduction complex, signaling through the TCR depends on a complex of proteins referred to collectively as CD3 (Figure 3-30). The CD3 complex is made up of three dimers: a ␦⑀ (delta epsilon) pair, a ␥⑀ (gamma epsilon) pair, and a third pair that is made up either of two CD3␨ (zeta) (a) molecules or a  (zeta, eta) heterodimer. (Note that the CD3 ␥ and ␦ chains are different from the chains that make up the ␥␦ TCR.) Like Ig␣ and Ig␤, the cytoplasmic tails of the CD3 molecules are studded with ITAM sequences that serve as docking sites for adapter proteins following activationinduced tyrosine phosphorylation. Each of the CD3 dimers contains negatively charged amino acids in its transmembrane domain that form ionic bonds with the positively charged residues on the intramembrane regions of the T-cell receptor. TCR NH2 23 222 S S 255 9 + + − − 30 NH2 NH2 90 + COOH COOH (248) (282) NH2 105 105 NH2 S S 201 εδ S S 184 140 S S S S 134 S S γε 91 S S 90 ζ ζ β S S 22 S S NH2 S S α 80 − − − − 116 130 130 106 160 COOH ITAM 185 COOH 185 COOH 150 COOH 143 COOH COOH TCR–CD3 (b) TCR ␥␧ ␦␧ ⫺ ⫹ ⫹ ⫹ ⫺ ␨␨ ⫺ ITAM FIGURE 3-30 Schematic diagram of the TCR-CD3 complex, which constitutes the T-cell antigen-binding receptor. (a) Components of the CD3 complex include the ␨␨ homodimers (alternately, a ␨␩ heterodimer) plus ␥⑀ and ␦⑀ heterodimers. The external domains of the ␥, ␦, and ⑀ chains of CD3 consist of immunoglobulin folds, which facilitates their interaction with the T-cell receptor and with each other. The long cytoplasmic tails of the CD3 chains contain a common sequence, the Immunoreceptor Tyrosine-based Activation Motif (ITAM), which functions in signal transduction; these sequences are shown as blue boxes. (b) Ionic interactions also may occur between the oppositely charged transmembrane regions in the TCR and CD3 chains. Proposed interactions among the CD3 components and the ␣␤ TCR are shown. [Adapted from M. E. Call and K. W. Wucherpfenning, 2004. Molecular mechanisms for the assembly of the T cell receptor-CD3 complex. Molecular Immunology 40:1295.] | Receptors and Signaling: B and T-Cell Receptors TABLE 3-4 CHAPTER 3 99 Selected T cell accessory molecules participating in T-cell signal transduction FUNCTION Name Ligand Adhesion Signal transduction Member of Ig superfamily CD4 Class II MHC ⫹ ⫹ ⫹ CD8 Class I MHC ⫹ ⫹ ⫹ CD2 (LFA-2) CD58 (LFA-3) ⫹ ⫹ ⫹ CD28 CD80, CD86 ? ⫹ ⫹ CTLA-4 CD80, CD86 ? ⫹ ⫺ CD45R CD22 ⫹ ⫹ ⫹ CD5 CD72 ? ⫹ ⫺ β S S D2 CD8 α S S D1 S S CD4 S S The T-cell receptor is noncovalently associated with a number of accessory molecules on the cell surface (Table 3-4). However, the only two such molecules that also recognize the MHC-peptide antigen are CD4 and CD8. Recall that mature T cells can be subdivided into two populations according to their expression of CD4 or CD8 on the plasma membrane. CD4⫹ T cells recognize peptides that are combined with class II MHC molecules, and function primarily as helper or regulatory T cells, whereas CD8⫹ T cells recognize antigen that is expressed on the surface of class I MHC molecules, and function mainly as cytotoxic T cells. CD4 is a 55 kDa monomeric membrane glycoprotein that contains four extracellular immunoglobulin-like domains (D1–D4), a hydrophobic transmembrane region, and a long cytoplasmic tail containing three serine residues that can be phosphorylated (Figure 3-31). CD8 takes the form of a disulfide-linked ␣␤ heterodimer or ␣␣ homodimer. (These are not the same as the ␣ and ␤ chains that constitute the TCR heterodimer.) Both the ␣ and ␤ chains of CD8 are small glycoproteins of approximately 30 to 38 kDa. Each chain consists of a single, extracellular, immunoglobulin-like domain, a stalk region, a hydrophobic transmembrane region, and a cytoplasmic tail containing 25 to 27 residues, several of which can be phosphorylated. The extracellular domains of CD4 and CD8 bind to conserved regions of MHC class II and MHC class I molecules respectively (see Figure 3-7b). The co-engagement of a single MHC molecule by both the TCR and its CD4 or CD8 coreceptor enhances the avidity of T-cell binding to its target. This co-engagement also brings the cytoplasmic domains of the TCR/CD3 and the respective co-receptor into close proximity, and it helps to initiate the cascade of intracellular events that activate a T cell. Signaling through the antigen receptor, even when combined with that through CD4 or CD8, is insufficient to activate a T cell that has had no prior contact with antigen (a naïve T cell). A naïve T cell needs to be simultaneously signaled through the TCR and its co-receptor, CD28, in order to be activated. The TCR and CD28 molecules on a naïve T cell must co-engage the MHC-presented peptide and the CD28 ligand, CD80 (or CD86), respectively, on the antigenpresenting cell for full activation to occur. The signaling events mediated through CD28, which include the stimulation of interleukin 2 synthesis by the T cell, were alluded to above and are discussed fully in Chapter 11. D3 D4 S S The T Cell Co-receptors CD4 and CD8 Also Bind the MHC S S FIGURE 3-31 General structure of the CD4 and CD8 coreceptors; the Ig-like domains are shown as circles. CD8 takes the form of an ␣␤ heterodimer or an ␣␣ homodimer. The monomeric CD4 molecule contains four Ig-fold domains; each chain in the CD8 molecule contains one. PA R T I 100 | Introduction CD45 TCR/CD3 CD4 CD28 ss PIP3 PIP3 Itk GADS P LAT SLP76 P P P VAV P P PLCγ1 P P P ZAP-70 P P P P Lck P Grb-2 SOS Ras DAG Ca2+ Akt PDK1 Grb-2 PI3K PKCθ Bax Bad Rac/Rho/ cdc42 Cytoskeletal reorganization Survival MAP kinase cascade Calmodulin Calcineurin Cytoplasm Nucleus NFAT NF-κB AP-1 Gene activation Lck is the First Tyrosine Kinase Activated in T Cell Signaling When the T-cell receptor interacts with its cell-bound antigen, receptors, co-receptors, and signaling molecules cluster into the cholesterol-rich lipid rafts of the plasma membrane (Figure 3-32). The Src-family tyrosine kinase Lck is normally found associated with CD4 and CD8, and the association between Lck and CD4 is particularly close. Antigen-induced clustering of the receptor–co-receptor complex brings Lck into the vicinity of the membrane-associated tyrosine phosphatase, CD45, which removes the inhibitory phosphate group on Lck. Reciprocal phosphorylation by nearby Lck molecules at their activating tyrosine sites (see Figure 3-9) then induces Lck to phosphorylate CD3 ITAM residues. (Note that these early events parallel those induced in B cells by the Src-family kinase Lyn, which is also regulated by the phosphatase activity of CD45.) Once the CD3 ITAMs are phosphorylated, a second tyrosine kinase, ZAP-70, docks at the phosphorylated tyrosine FIGURE 3-32 Signal transduction pathways emanating from the TCR. T cell antigen binding activates the src-family kinase Lck, which phosphorylates the kinase ZAP-70. ZAP-70 in turn phosphorylates the adapter molecules LAT, SLP76, and GADS which form a scaffold enabling the phosphorylation and activation of PLC␥1 and PKC␪ with the consequent effects on transcription factor activation described in the text. The GEF proteins Vav and SOS are also activated on binding to LAT, leading to activation of the Ras/MAP kinase transcription factor pathway and the Rac/Rho/cdc42 pathway, leading to changes in cell shape and motility. PI3 kinase, translocated to the cytoplasmic side of CD28, forms PIP3, inducing localization of the enzymes PDK1 and Akt to the membrane. This leads to further NF-␬B activation and increased cell survival as described. residues of the CD3 chains. ZAP-70 is activated by Lckmediated phosphorylation and goes on to phosphorylate many adapter molecules including SLP-76 and LAT, as well as enzymes important in T-cell activation, such as PLC␥1. T Cells Use Downstream Signaling Strategies Similar to Those of B Cells Just as in B cells, signals initiating at the antigen receptor of T cells with tyrosine phosphorylation events are then fanned out to intracellular enzymes and transcription factors using a network of adapter molecules and enzymes. In T cells, one of the earliest adapter molecules to be incorporated into the signaling complex is LAT (Linker protein of Activated T cells), a transmembrane protein associated with lipid rafts in the plasma membrane. Following TCR ligation, LAT is phosphorylated on multiple residues by ZAP-70, and these phosphorylated residues now provide docking sites for several important enzymes bearing SH2 domains, including PLC␥1 (see Figure 3-32). Phosphorylated LAT also binds to Receptors and Signaling: B and T-Cell Receptors the adapter protein GADS, which is constitutively associated with the adapter SLP-76. This combination of adapter proteins is critical to T-cell receptor signaling, providing the structural framework for most downstream signaling events. Many of those downstream events will now be familiar. PLC␥1, localized to the plasma membrane by binding to LAT, is further activated by tyrosine phosphorylation, mediated by the kinase Itk (which belongs to a family of kinases referred to as Tec kinases). As described earlier, PLC␥1 breaks down PIP2, releasing IP3, which induces the release of calcium and the activation of NFAT via calcineurin activation. The DAG created by PIP2 hydrolysis binds, in T cells, to a specialized form of PKC called PKC␪ (theta). As described above, this part of the signaling cascade similarly culminates in the degradation of the inhibitors of NF-␬B and the translocation of the active transcription factor into the nucleus (see Figure 3-17). Phosphorylated LAT also associates with the SH2 domain of Grb2, the now-familiar adapter molecule that brings in components of the Ras pathway to the signaling complex. Recall that Grb2 binds constitutively to SOS, the GEF that | CHAPTER 3 101 facilitates activation of the Ras pathway. In T cells, the Ras pathway is important both to the activation of the transcription factor AP-1, which functions to signal cytokine secretion, and to the passage of the signals that reorganize the actin cytoskeleton for directed cytokine release. Thus, as for B cells, TCR-antigen binding leads to a multitude of consequences including transcription factor upregulation, reorganization of the cytoskeleton, and cytokine secretion. Again, as for B cells, T-cell signaling also affects the expression of adhesion molecules such as integrins on the cell surface, and chemokines, which has subsequent effects on cell localization. Clearly the description of adaptive immune signaling offered in this chapter represents just the tip of the iceberg, and these signaling cascades contain more components and outcomes that can be alluded to in this brief outline. However, just as the adapter proteins provide a scaffold for the immune system to organize its signaling proteins, so we hope that this chapter has provided a similar scaffold for the organization of the reader’s thoughts regarding these fascinating and complex processes. S U M M A R Y ■ ■ ■ ■ ■ ■ ■ Antigens bind to receptors via noncovalent bonding interactions. The interactions between antigens and receptors of the immune system are enhanced by simultaneous interactions between lymphocyte-expressed co-receptors and molecules on antigen-presenting cells or on complex antigens. Most receptor-antigen interactions are multivalent, and this multivalency significantly increases the avidity of the receptor-antigen binding interaction. Binding of antigen to receptor induces a signaling cascade in the receptor-bearing cell, which leads to alterations in the motility, adhesive properties, and transcriptional program of the activated cell. Antigen signaling is initiated in both T and B cells by antigen-mediated receptor clustering. The clustered receptors are located in specialized regions of the membranes called lipid rafts. The CD3 and Ig␣/Ig␤ proteins, which are T- and B-cell receptor-associated signal transduction elements, are phosphorylated on Immunoreceptor Tyrosine Activation Motifs (ITAMs), and these serve as docking points for adapter molecules. Downstream signaling enzymes and GEFs dock onto the adapter molecules and make contact with their substrates. These enzymes include phospholipase C␥, which breaks down PIP2 into DAG and IP3. IP3 interaction with ERlocated receptors leads to release of intracellular calcium and activation of calcium-regulated proteins such as calcineurin phosphatase. Calcineurin dephosphorylates the transcription factor NFAT, allowing it access to the ■ ■ ■ ■ ■ ■ ■ nucleus. DAG binds and activates protein kinase C, leading eventually to NF-␬B activation. Docking of the adapter molecule Grb2 onto adapter proteins facilitates its binding to the GEF protein SOS and activation of the MAP kinase pathway, resulting in activation of the transcription factor AP-1. The antigen receptor on B cells is a membrane-bound form of the four-chain immunoglobulin molecule that the B cell secretes upon stimulation. An immunoglobulin molecule is commonly known as an antibody. Antibodies have two heavy and two light chains. The antibodies secreted by B cells upon stimulation are classified according to the amino acid sequence of the heavy chain, and antibodies of different classes perform different functions during an immune response. Antigen signaling in B cells proceeds according to signaling strategies shared among many cell types. The antigen receptor on T cells is not an immunoglobulin molecule, although its protein domains are classified as belonging to the immunoglobulin superfamily of proteins. Most T cells bear receptors made up of an ␣␤ heterodimer that recognizes a complex antigen, made up of a short peptide inserted into a groove on the surface of a protein encoded by the major histocompatibility complex (MHC) of genes. Some T cells bear receptors made up of ␥␦ receptor chains that recognize a different array of antigens. Antigen signaling in T cells shares many characteristic strategies with B cell signaling. 102 PA R T I | Introduction R E F E R E N C E S Andreotti, A. H., P. L. Schwartzberg, R. E. Joseph, and L. J. Berg. 2010. T-cell signaling regulated by the Tec family kinase, Itk. Cold Spring Harbor Perspectives in Biology 2:a002287. Porter, R. R. (1972). Lecture for the Nobel Prize for physiology or medicine 1972: Structural studies of immunoglobulins. Scandinavian Journal of Immunology 34(4):381–389. Chaturvedi, A., Z. Siddiqui, F. Bayiroglu, and K. V. Rao. 2002. A GPI-linked isoform of the IgD receptor regulates resting B cell activation. Nature Immunology 3:951. Tiselius, A. 1937. Electrophoresis of serum globulin. I. The Biochemical Journal 31:313. Choudhuri, K., and M. L. Dustin. 2010. Signaling microdomains in T cells. FEBS Letters 584:4823. Edelman G. M., B. A. Cunningham, W. E. Gall, P. D. Gottlieb, U. Rutishauser, and M. J. Waxdal. 1969. The covalent structure of an entire gammaG immunoglobulin molecule. Proceedings of the National Academy of Sciences U.S.A. 63:78. Finlay, D., and D. Cantrell. The coordination of T-cell function by serine/threonine kinases. 2011. Cold Spring Harbor Perspectives in Biology 3:a002261. Fitzgerald, K. A., and Z. J. Chen. 2006. Sorting out Toll signals. Cell 125:834. Heidelberger, M., and F. E. Kendall. 1936. Quantitative studies on antibody purification : I. The dissociation of precipitates formed by Pneumococcus specific polysaccharides and homologous antibodies. Journal of Experimental Medicine 64:161. Heidelberger, M., and K. O. Pedersen. 1937. The Molecular Weight of Antibodies. Journal of Experimental Medicine 65:393. Kanneganti, T. D., M. Lamkanfi, and G. Nunez. 2007. Intracellular NOD-like receptors in host defense and disease. Immunity 27:549. Koretzky, G. A., and P. S. Myung. 2001. Positive and negative regulation of T-cell activation by adaptor proteins. Nature Reviews Immunology 1:95. Kurosaki, T. 2011. Regulation of BCR signaling. Molecular Immunology 48:1287. Love, P. E., and S. M. Hayes. 2010. ITAM-mediated signaling by the T-cell antigen receptor. Cold Spring Harbor Perspectives in Biology 2:a002485. Macian, F. 2005. NFAT proteins: Key regulators of T-cell development and function. Nature Reviews Immunology 5:472. Nisonoff, A., F. C. Wissler, and L. N. Lipman. 1960. Properties of the major component of a peptic digest of rabbit antibody. Science 132:1770. Palacios, E. H., and A. Weiss. 2004. Function of the Src-family kinases, Lck and Fyn, in T-cell development and activation. Oncogene 23:7990. Park, S. G., J. Schulze-Luehrman, M. S. Hayden, N. Hashimoto, W. Ogawa, M. Kasuga, and S. Ghos. 2009. The kinase PDK1 integrates T cell antigen receptor and CD28 coreceptor signaling to induce NF-kappaB and activate T cells. Nature Immunology 10:158. Tiselius, A., and E. A. Kabat. 1938. Electrophoresis of immune serum. Science 87:416. Tiselius, A., and E. A. Kabat. 1939. An electrophoretic study of immune sera and purified antibody preparations. Journal of Experimental Medicine 69:119. Wang, H., T. A. Kadlecek, B. B. Au-Yeung, H. E. Goodfellow, L. Y. Hsu, T. S. Freedman, and A. Weiss. 2010. ZAP-70: An essential kinase in T-cell signaling. Cold Spring Harbor Perspectives in Biology 2:a002279. Yamamoto, M., and K. Takeda. 2010. Current views of toll-like receptor signaling pathways. Gastroenterology Research and Practice 2010:240365. Useful Web Sites www.genego.com This site offers a searchable database of metabolic and regulatory pathways. www.nature.com/subject/cellsignaling This is a collection of original research articles, reviews and commentaries pertaining to cell signaling. http://stke.sciencemag.org This is the Signal Transduction Knowledge Environment Web site, maintained by Science magazine. It is an excellent site, but parts of it are closed to those lacking a subscription. www.signaling-gateway.org This gateway is powered by University of California at San Diego and is supported by Genentech and Nature. An excellent resource, complete with featured articles. www.qiagen.com/geneglobe/pathwayview. aspx Qiagen. A useful commercial Web site. www.biosignaling.com Part of Springer Science⫹Business Media. www.youtube.com Many excellent videos and animations of signaling Web sites are available on YouTube, too numerous to mention here. Just type your pathways into a Web browser and go. But do be cognizant of the derivation of your video. Not all videos are accurate, so check your facts with the published literature. www.hhmi.org/biointeractive/immunology/tcell. html Howard Hughes Medical Institute (HHMI) movie: Cloning an Army of T Cells for Immune Defense. Receptors and Signaling: B and T-Cell Receptors S T U D Y | CHAPTER 3 103 Q U E S T I O N S 1. The NFAT family is a ubiquitous family of transcription factors. a. Under resting conditions, where is NFAT localized in a cell? b. Under activated conditions, where is NFAT localized in a cell? c. How is it released from its resting condition and per- mitted to relocalize? d. Immunosuppressant drugs such as cyclosporin act via inhibition of the calcineurin phosphatase. If NFAT is ubiquitous, how do you think these drugs might act with so few side effects on other signaling processes within the body? 2. In the early days of experiments designed to detect the T-cell receptor, several different research groups found that antibodies directed against immunoglobulin proteins appeared to bind to the T-cell receptor. Given what you know about the structure of immunoglobulins and the T-cell receptor, why is this not completely surprising? 3. True or false? Explain your answers. Interactions between receptors and ligands at the cell surface: a. are mediated by covalent interactions. b. can result in the creation of new covalent interactions within the cell. 4. Describe how the following experimental manipulations were used to determine antibody structure. a. Reduction and alkylation of the antibody molecule b. Enzymatic digestion of the antibody molecule c. Antibody detection of immunoglobulin fragments 5. What is an ITAM, and what proteins modify the ITAMs in Ig␣ and Ig␤? 6. Define an adapter protein. Describe how an interaction between proteins bearing SH2 and phosphorylated tyrosine (pY) groups helps to transduce a signal from the T-cell receptor to the ZAP-70 protein kinase. 7. IgM has ten antigen-binding sites per molecule, whereas IgG only has two. Would you expect IgM to be able to bind five times as many antigenic sites on a multivalent antigen as IgG? Why/why not? 8. You and another student are studying a cytokine receptor on a B cell that has a Kd of 10⫺6 M. You know that the cytokine receptor sites on the cell surface must be at least 50% occupied for the B cell to receive a cytokine signal from a helper T cell. Your lab partner measures the cytokine concentration in the blood of the experimental animal and detects a concentration of 10⫺7 M. She tells you that the effect you have been measuring could not possibly result from the cytokine you’re studying. You disagree. Why? 9. Activation of Src-family kinases is the first step in several different types of signaling pathways. It therefore makes biological sense that the activity of this family of tyrosine kinases is regulated extremely tightly. Describe how phosphorylation of Src-family kinases can deliver both activating and inhibitory signals to Src kinases. 10. You have generated a T-cell clone in which the Src-family tyrosine kinase Lck is inactive. You stimulate that clone with its cognate antigenic peptide, presented on the appropriate MHC platform and test for interleukin 2 secretion, as a measure of T cell activation. Do you expect to see IL-2 secretion or not? Explain. 11. Name one protein shown to be defective in many cases of X-linked agammaglobulinemia, and describe how a reduction in the activity of this protein could lead to immunodeficiency. 12. The B- and T-cell receptor proteins have remarkably short intracytoplasmic regions of just a few amino acids. How can you reconcile this structural feature with the need to signal the presence of bound antigen to the interior of the cell? 13. Describe one way in which the structure of antibodies is superbly adapted to their function. 14. As a graduate student, your adviser has handed you a T-cell clone that appears to be constitutively (always) activated, although at a low level, even in the absence of antigenic stimulation, and he has asked you to figure out why. Your benchmate suggests you start by checking out the sequence of its lck gene, or the status of the Csk activity in the cell. You agree that those are good ideas. What is your reasoning? This page lelt intentionally blank. Receptors and Signaling: Cytokines and Chemokines Immunoglobulin family receptors Hematopoietin-type receptors (class I) Interferon-type receptors (class II) S S α β JAK JAK S S T he hundreds of millions of cells that comprise the vertebrate immune system are distributed throughout the body of the host (see Chapter 2). Some cells circulate through the blood and lymph systems, whereas others are sessile (remain in place) in the primary and secondary lymphoid tissues, the skin, and the mucosa of the respiratory, alimentary, and genito-urinary tracts. The key to success for such a widely dispersed organ system is the ability of its various components to communicate quickly and efficiently with one another, so that the right cells can home to the appropriate locations and take the necessary measures to destroy invading pathogens. Molecules that communicate among cells of the immune system are referred to as cytokines. In general, cytokines are soluble molecules, although some also exist in membrane-bound forms. The interaction of a cytokine with its receptor on a target cell can cause changes in the expression of adhesion molecules and chemokine receptors on the target membrane, thus allowing it to move from one location to another. Cytokines can also signal an immune cell to increase or decrease the activity of particular enzymes or to change its transcriptional program, thereby altering and enhancing its effector functions. Finally, they can instruct a cell when to survive and when to die. In an early attempt to classify cytokines, immunologists began numbering them in the order of their discovery, and naming them interleukins. This name reflects the fact that interleukins communicate between (Latin, inter) white blood cells (leukocytes). Examples include interleukin 1 (IL-1), secreted by macrophages, and interleukin 2 (IL-2), secreted by activated T cells. However, many cytokines that were named prior to this attempt at rationalizing nomenclature have resisted reclassification, and so students will come across cytokines such as Tumor Necrosis Factor or Interferons, that are also “interleukins” in all but name. Although the term cytokine refers to all molecules that communicate among immune cells, the name chemokine is used specifically to describe that subpopulation of 4 S S α γ β TNF receptors IL-17 receptors Chemokine receptors G protein Cytokine and Chemokine Receptor Families ■ ■ General Properties of Cytokines and Chemokines Six Families of Cytokines and Associated Receptor Molecules ■ Cytokine Antagonists ■ Cytokine-Related Diseases ■ Cytokine-Based Therapies cytokines that share the specific purpose of mobilizing immune cells from one organ, or indeed, from one part of an organ, to another. Chemokines belong to the class of molecules called chemoattractants, molecules that attract cells by influencing the assembly, disassembly, and contractility of cytoskeleton proteins and the expression of cell-surface adhesion molecules. Chemokines attract cells with the appropriate chemokine receptors to regions where the chemokine concentration is highest. For example, chemokines are important in attracting cells of 105 106 PA R T I | Introduction the innate immune system to the site of infection and inducing T cells to move toward antigen-presenting cells in the secondary lymphoid tissues. Leukocytes change their pattern of expression of chemokine receptors over the course of an immune response, first migrating to the secondary immune organs, in which they undergo differentiation to mature effector cells, and then moving out into the affected tissues to fight the infection, responding to different chemokine gradients with each movement. As we will learn in a later section, chemokines are also capable of instructing cells to alter their transcriptional programs. The classification and nomenclature of chemokines is more logical than that of interleukins, and is based on their biochemical structures. Although chemokines technically fall under the umbrella classification of “cytokines,” normal usage is evolving such that the term chemokine is used when referring to molecules that move immune cells from place to place, and the term cytokine is employed when referring to any other messenger molecule of the immune system. Like all signaling molecules, cytokines can be further classified on the basis of the distance between the cell secreting the signaling ligand and the cell receiving that chemical signal. Cytokines that act on cells some distance away from the secreting cell, such that they must pass through the bloodstream before reaching their target, are referred to as endocrine (Figure 4-1). Those that act on cells near the secreting cell, such that the cytokine merely has to diffuse a few Ångstroms through tissue fluids or Circulation Endocrine action Paracrine action Distant cell Nearby cell Autocrine action FIGURE 4-1 Most immune system cytokines exhibit autocrine and/or paracrine action; fewer exhibit endocrine action. across an immunological synapse, are referred to as paracrine. Sometimes, a cell needs to receive a signal through its own membrane receptors from a cytokine that it, itself, has secreted. This type of signaling is referred to as autocrine. Of note, the T-cell interleukin IL-2 acts effectively in all three modes. Unlike the classical hormones, such as insulin and glucagon, that generally act at long range in an endocrine fashion, many cytokines act over a short distance in an autocrine or paracrine fashion. We begin this chapter with an introduction to the general properties of cytokines and chemokines followed by a discussion of the specific receptors and signaling pathways used by the six families of immune system cytokines and chemokines. Next, we describe the ways in which cytokine signaling can be regulated by antagonists. Finally, we turn to the role of cytokines and chemokines in disease and medicine. General Properties of Cytokines and Chemokines The activity of cytokines was first recognized in the mid1960s, when supernatants derived from in vitro cultures of lymphocytes were found to contain soluble factors, usually proteins or glycoproteins, that could regulate proliferation, differentiation, and maturation of immune system cells. Production of these factors by cultured lymphocytes was induced by activation with antigens or with nonspecific mitogens (molecules inducing cell division, or mitosis). However, biochemical isolation and purification of cytokines was initially hampered because of their low concentrations in the culture supernatants and the absence of well-defined assay systems for individual cytokines. The advent of hybridoma technology (see Chapter 20) allowed the production of artificially generated T-cell tumors that constitutively produced IL-2, allowing for its purification and characterization. Gene cloning techniques developed during the 1970s and 1980s then made it possible to generate pure cytokines by expressing the proteins from cloned genes derived from hybridomas or from normal leukocytes, after transfection into bacterial or yeast cells. Using these pure cytokine preparations, researchers were able to identify cell lines whose growth depended on the presence of a particular cytokine, thus providing them with biological cytokine assay systems. Since then, monoclonal antibodies specific for many cytokines have made it possible to develop rapid, quantitative, cytokine-specific immunoassays. ELISA assays measure the concentrations of cytokines in solution, Elispot assays quantitate the cytokines secreted by individual cells, and cytokine-specific antibodies can be used to identify cytokine-secreting cells using intracellular cytokine staining followed by flow cytometry or immuno-fluorescence microscopy (see Chapter 20). Receptors and Signaling: Cytokines and Chemokines Cytokines Mediate the Activation, Proliferation, and Differentiation of Target Cells Cytokines bind to specific receptors on the membranes of target cells, triggering signal transduction pathways that ultimately alter enzyme activity and gene expression (Figure 4-2). The susceptibility of a target cell to a particular cytokine is determined by the presence of specific membrane receptors. In general, cytokines and their fully assembled receptors exhibit very high affinity for one another, with dissociation constants for cytokines and their receptors ranging from 108 to 1012 M1. Because their receptor affinities are so high and because cytokines are often secreted in close proximity to their receptors, such that the cytokine concentration is not diluted by diffusion (as mentioned in Chapter 3), the secretion of very few cytokine molecules can mediate powerful biological effects. Cytokines regulate the intensity and duration of the immune response by stimulating or inhibiting the activation, proliferation, and/or differentiation of various cells, by regulating the secretion of other cytokines or of antibodies, or in some cases by actually inducing programmed cell death in the target cell. In addition, cytokines can modulate the Inducing stimulus Cytokine gene activation Cytokine-producing cell Cytokine secretion Cytokine receptor Signal Gene or enzyme activation Target cell Biological effect (e.g., proliferation, differentiation, cell death) FIGURE 4-2 Overview of the induction and function of cytokines. An inducing stimulus, which may be an antigen or another cytokine, interacts with a receptor on one cell, inducing it to secrete cytokines that in turn act on receptors of a second cell, bringing about a biological consequence. In the case of IL-2, both cells may be antigen-activated T cells that secrete IL-2, which acts both on the secreting cell and on neighboring, activated T cells. | CHAPTER 4 107 expression of various cell-surface receptors for chemokines, other cytokines, or even for themselves. Thus, the cytokines secreted by even a small number of antigen-activated lymphocytes can influence the activity of many different types of cells involved in the immune response. Cytokines exhibit the attributes of pleiotropy, redundancy, synergism, antagonism, and cascade induction (Figure 4-3), which permit them to regulate cellular activity in a coordinated, interactive way. A cytokine that induces different biological effects depending on the nature of the target cells is said to have a pleiotropic action, whereas two or more cytokines that mediate similar functions are said to be redundant. Cytokine synergy occurs when the combined effect of two cytokines on cellular activity is greater than the additive effects of the individual cytokines. In some cases, the effects of one cytokine inhibit or antagonize the effects of another. Cascade induction occurs when the action of one cytokine on a target cell induces that cell to produce one or more additional cytokines. Cytokines Have Numerous Biological Functions Although a variety of cells can secrete cytokines that instruct the immune system, the principal producers are TH cells, dendritic cells, and macrophages. Cytokines released from these cell types are capable of activating entire networks of interacting cells (Figure 4-4). Among the numerous physiological responses that require cytokine involvement are the generation of cellular and humoral immune responses, the induction of the inflammatory response, the regulation of hematopoiesis, and wound healing. The total number of proteins with cytokine activity grows daily as research continues to uncover new ones. Table 4-1 summarizes the activities of some commonly encountered cytokines. An expanded list of cytokines can be found in Appendix II. Note, however, that many of the listed functions have been identified from analyses of the effects of recombinant cytokines, sometimes added alone to in vitro systems at nonphysiologic concentrations. In vivo, cytokines rarely, if ever, act alone. Instead, a target cell is exposed to a milieu containing a mixture of cytokines whose combined synergistic or antagonistic effects can have a wide variety of consequences. In addition, as we have learned, cytokines often induce the synthesis of other cytokines, resulting in cascades of activity. Cytokines Can Elicit and Support the Activation of Specific T-Cell Subpopulations As described in Chapters 2 and 11, helper T cells can be classified into subpopulations, each of which is responsible for the support of a different set of immune functions. For example, TH1 cells secrete cytokines that promote the differentiation and activity of macrophages and cytotoxic T cells, thus leading to a primarily cytotoxic immune response, in which cells that have been infected with viruses and 108 PA R T I | Introduction Target Cell Effect CASCADE INDUCTION Activation Proliferation Differentiation Activated TH cells PLEIOTROPY B cell IL-4 Proliferation and differentiation IFN-γ Cytotoxic T cell precursor Activated TH cells Proliferation Mast cell Macrophage REDUNDANCY IL-2 IL-4 IL-5 Proliferation IL-12 Activated TH cells B cell SYNERGY IL-4 + IL-5 Activated TH cells Induces class switch to IgE B cell Activated TH cells ANTAGONISM Blocks class switch to IgE induced by IL-4 IL-4 IFN-γ Activated TH cells B cell IFN-γ, TNF, IL-2, and other cytokines FIGURE 4-3 Cytokine attributes of (a) pleiotropy, redundancy, synergism, antagonism, and (b) cascade induction. intracellular bacteria are recognized and destroyed. The cytokines IL-12 and interferon (IFN)  induce TH1 differentiation. In contrast, TH2 cells activate B cells to make antibodies, which neutralize and bind extracellular pathogens, rendering them susceptible to phagocytosis and complement-mediated lysis. IL-4 and IL-5 support the generation of TH2 cells. TH17 cells promote the differentiation of activated macrophages and neutrophils, and support the inflammatory state; their generation is induced by IL-17 and IL-23. The differentiation and activity of each distinctive T-cell subpopulation is therefore supported by the binding of different combinations of cytokines to T-cell surface receptors, with each cytokine combination inducing its own characteristic array of intracellular signals, and sending the helper T cell down a particular differentiation pathway. | Receptors and Signaling: Cytokines and Chemokines CHAPTER 4 109 IFN-α IFN-β IFN-γ IL-6 IL-10 IL-15 Macrophage IL-1 IFN-α IL-6 IFN-γ IL-8 IFN-α IL-10 IL-12 IL-1 IL-15 IFN-α IL-18 IFN-β IL-1 TNF-α IL-3 TNF-β IL-4 IL-10 IL-13 IL-6 IFN-α IL-10 TNF-β IL-10 TNF-β TNF-α IL-1 IL-8 TNF-α IL-1 IL-8 IFN-α IFN-β IFN-γ TNF-α TNF-β Neutrophil TNF-α IL-1 IFN-γ IL-3 TNF-α IL-4 TNF-β IL-8 B-cell IL-1 IFN-α IL-2 IFN-β IL-4 IFN-γ IL-6 IL-10 IL-13 IL-14 IL-1 IL-3 IL-4 IL-8 Basophil IL-6 IL-10 IL-2 IL-4 IL-6 IL-8 IL-9 IL-16 IL-17 IFN-α IFN-β IFN-γ T-cell IL-3 IL-4 IL-5 Eosinophil IL-3 IL-9 IL-10 IL-4 IL-4 IL-5 IL-4 IL-10 IL-4 IL-4 Mast cell FIGURE 4-4 The cells of the immune system are subject to control by a network of cytokine actions. Cell Activation May Alter the Expression of Receptors and Adhesion Molecules The ability of cytokines to activate most, if not all, members of particular immune cell subpopulations appears to conflict with the established specificity of the immune system. What keeps cytokines from activating all T cells, for example, in a nonspecific fashion during the immune response? In order for a cell to respond to a signaling molecule, it must express receptors for that molecule, and responsiveness to a molecular signal can thus be controlled by signal receptor expression. For example, antigen stimulation of a T cell induces alterations in the T-cell surface expression of chemokine receptors. Reception of chemokine signals through these receptors therefore instructs only those cells that have previously been activated by antigen to migrate to nearby lymph nodes or to the spleen. Furthermore, activationinduced changes in the adhesion molecules that are expressed on the cell membrane ensures that stimulated cells migrate to, and then remain in, the location best suited to their function. T-cell activation by antigen also up-regulates the expression of the receptors for cytokines that provide proliferative signals, such as IL-2 (as described in Chapter 3), and also for differentiative cytokines such as IL-4. In this way, following antigen encounter, only those T cells that have been activated by antigen are primed to relocate and to receive the proliferative and differentiative signals they need to function as a mature immune effector cell. This 110 PA R T I TABLE 4-1 | Introduction Functional groups of selected cytokines* Cytokine Secreted by† Targets and effects SOME CY TOKINES OF INNATE IMMUNITY Interleukin 1 (IL-1) Monocytes, macrophages, endothelial cells, epithelial cells Vasculature (inflammation); hypothalamus (fever); liver (induction of acute phase proteins) Tumor necrosis factor- (TNF-) Macrophages, monocytes, neutrophils, activated T cells and NK cells Vasculature (inflammation); liver (induction of acute phase proteins); loss of muscle, body fat (cachexia); induction of death in many cell types; neutrophil activation Interleukin 12 (IL-12) Macrophages, dendritic cells NK cells; influences adaptive immunity (promotes TH1 subset) Interleukin 6 (IL-6) Macrophages, endothelial cells, and TH2 cells Liver (induces acute phase proteins); influences adaptive immunity (proliferation and antibody secretion of B-cell lineage) Interferon- (IFN-) (this is a family of molecules) Macrophages dendritic cells, virus-infected cells Induces an antiviral state in most nucleated cells; increases MHC Class I expression; activates NK cells Interferon  (IFN-) Macrophages, dendritic cells, virus-infected cells Induces an antiviral state in most nucleated cells; increases MHC Class I expression; activates NK cells SOME CY TOKINES OF ADAPTIVE IMMUNITY Interleukin 2 (IL-2) T cells T-cell proliferation; can promote AICD. NK cell activation and proliferation; B-cell proliferation Interleukin 4 (IL-4) TH2 cells, mast cells Promotes TH2 differentiation; isotype switch to IgE Interleukin 5 (IL-5) TH2 cells Eosinophil activation and generation Transforming growth factor  (TGF-) T cells, macrophages, other cell types Inhibits T-cell proliferation and effector functions; inhibits B-cell proliferation; promotes isotype switch to IgA; inhibits macrophages Interferon  (IFN-) TH1 cells, CD8 cells, NK cells Activates macrophages; increases expression MHC Class I and Class II molecules; increases antigen presentation *Many cytokines play roles in more than one functional category. † Only the major cell types providing cytokines for the indicated activity are listed; other cell types may also have the capacity to synthesize the given cytokine. Activated cells generally secrete greater amounts of cytokine than unactivated cells. pattern of activation-induced alteration in the cell surface expression of adhesion molecules, chemokine receptors, and cytokine receptors is a common strategy employed by the immune system. Cytokines Are Concentrated Between Secreting and Target Cells During the process of T-cell activation by an antigen-presenting dendritic cell, or of B-cell activation by a cognate T cell, the respective pairs of cells are held in close juxtaposition for many hours (see Chapter 14). Over that time period, the cells release cytokines that bind to relevant receptors on the partner cell surface, without ever entering the general circulation. Furthermore, during this period of close cell-cell contact, the secretory apparatus of the stimulating cell is oriented so that the cytokines are released right at the region of the cell membrane that is in closest contact with the recipient cell (see Figure 3-4). The close nature of the cell-cell interaction and the directional release of cytokines by the secretory apparatus means that the effective concentration of cytokines in the region of the membrane receptors may be orders of magnitude higher than that experienced outside the contact region of the two cells. Thus, discussions of membrane receptor affinity and cytokine concentrations within tissue fluids must always take into account the biology of the responding system and the geography of the cell interactions involved. In addition, the half-life of cytokines in the bloodstream or other extracellular fluids into which they are secreted is usually very short, ensuring that cytokines usually act for only a limited time and over a short distance. Signaling Through Multiple Receptors Can Fine Tune a Cellular Response Cytokine and chemokine signaling in the immune response can be a strikingly complex and occasionally redundant affair. Effector molecules such as cytokines can bind to more Receptors and Signaling: Cytokines and Chemokines than one receptor, and receptors can bind to more than one signaling molecule. Nowhere is the latter concept more clearly illustrated than in the chemokine system, in which approximately 20 receptors bind to close to 50 distinct chemokines (see Appendix III). Effector molecule signaling can also cooperate with signaling through antigen-specific receptors. Signals received through more than one receptor must then be integrated at the level of the biological response, with multiple pathways acting to tune up or tune down the expression of particular transcription factors or the activity of particular enzymes. Thus, the actual biological response mounted by a cell to a particular chemical signal depends not only on the nature of the individual receptor for that signal, but also on all of the downstream adapters and enzymes present in the recipient cell. | CHAPTER 4 111 Six Families of Cytokines and Associated Receptor Molecules In recent years, immunologists have enjoyed an explosion of information about new cytokines and cytokine receptors as a result of advances in genomic and proteomic analysis. Advances Box 4-1 describes a recently developed proteomic approach to the search for new, secreted cytokines and illustrates the manner in which a sophisticated appreciation of the molecular and cell biology of secretory pathways aids in the identification of new cytokines. The purpose of this chapter is not to provide an exhaustive list of cytokines and their receptors (see Appendices II and III for a comprehensive and current list of cytokines and chemokines), but rather to outline BOX 4-1 ADVANCES Methods Used to Map the Secretome The related approaches of genomics and proteomics provide scientists with tools they can use to assess the complex changes that occur in gene and protein expression induced by stimuli, such as antigen or cytokine stimulation. Vast arrays of information regarding the derivation and readout of genes in different cells and organisms, and the expression of particular proteins, can be analyzed and presented in ways not available to scientists just a few years ago. Recently, the science of proteomics has been extended to address the mapping of proteins that are secreted by various cell types. The array of proteins secreted by a cell is referred to as its secretome, and the secretome can be more formally defined as the “proteins released by a cell, tissue, or organisms through classical and nonclassical secretion mechanisms.” Scientists first became interested in the concept of the secretome as a way to diagnose and identify various types of cancer. They reasoned that they could use the set of proteins secreted into the serum or other tissue fluids as a biological marker for spe- cific tumor types. If particular proteins can be shown to be secreted at high concentration only under conditions of malignancy, then rapid and inexpensive tests can be developed that have the potential to screen for tumors at an early stage, when they are still amenable to treatment. Although such tumor-specific profiles of secreted proteins are surprisingly difficult to develop, given the range of mutations associated with the generation of a cancer, the ability to diagnose a tumor at an early stage using only a serum sample provides intense motivation, and many such attempts are ongoing. The approaches used to define a set of cancer secretomes have since been applied to studies of many other, nonmalignant cell populations for which the description of a secretome would be a useful analytical tool. These populations include stem cells, cells of the immune system, and adipose cells. Given the diversity of cytokines that can be secreted by a single cell, and the manner in which the activities of cytokines can interact at the level of the target cell, cytokine biology is a superb target for such a global approach. A recent secretome analysis (Botto et al., 2011) addressed the question of how the human cytomegalovirus induces the formation of new blood vessels (angiogenesis). Virus-free supernatant from virus-infected endothelial cells was found to induce angiogenesis. Secretome analysis of the infected endothelial cell supernatant revealed the presence of multiple cytokines, including IL-8, GM-CSF, and IL-6. The addition of a blocking anti-IL-6 antibody at the same time as the virus-free supernatant was then shown to inhibit its angiogenic activity, thus demonstrating that it was the IL-6 activity in the supernatant that was primarily responsible for inducing the new blood vessel growth. One difficulty that is frequently encountered in trying to analyze the secretome of a particular type of cell is the need of many cells to grow in a tissue culture fluid supplemented with serum, which is itself a complex mixture of proteins. In this case it is important to distinguish between proteins released by the cells under study and those which were originally present in the serum. Several (continued) PA R T I 112 ADVANCES | Introduction BOX 4-1 (continued) techniques are available to discriminate between secreted proteins and those from the tissue culture media, including adding inhibitors of secretion to some cultures and then comparing those proteins present in the culture supernatant in the presence and absence of inhibitors. Alternatively, culturing the cells in the presence of radioisotopes that only label newly synthesized proteins, such as 35S methionine, can be used to distinguish these proteins from preexisting proteins in the culture medium. In the case, such as that described above, that a cell line is being tested to determine whether it secretes a set of cytokines for which antibody assays already exist, two different types of multiplex measurements may be used (see Figure 1). Both of these approaches utilize antibodies to the array of cytokines to be analyzed, attached to some sort of solid phase support. This support may be glass, a membrane, or a set of beads, with each antibody attached to a bead of a different color. The sample of tissue culture fluid is added to the solid phase antibody, excess fluid is washed away, and then biotinylated antibodies are added. (Biotin, a small molecule, is used because it has an extremely high affinity for a protein, streptavidin, and is therefore used to couple two molecules together in assays such as these. For more details, see Chapter 20.) After antibody binding, the excess biotinylated antibodies are removed by washing and the cytokine concentrations are assessed by the addition of fluorescent streptavidin, which will bind to the biotin. A fluorescent signal indicates the presence of the cytokine in the sample, and the level of the signal reveals its concentration Since each bead fluoresces at a different wavelength, the fluorescence associated with each cytokine can be distinguished. Various bioinformatics tools have been developed that have particular application to secretome analysis. These include SignalP, which identifies the presence of signal peptides and also shows the location of signal peptide cleavage sites in bacterial and eukaryotic proteins. In addition, SecretomeP can be used for the pre- Planar arrays a-IL-8 a-TGF-β a-bFGF a-TNF-α Bead assays Panel of antibodies recognizing cytokines, chemokines and growth factors a-IL-8 a-bFGF a-TGF-β a-TNF-α Antibodies immobilized on beads that fluoresce at different wavelengths Antibodies printed on glass or membrane Incubation with sample Incubation with biotinylated antibodies a-IL-8 a-TGF-β a-bFGF a-TNF-α Incubation with labelled streptavidin Detection of signal, data analysis Fluorescent or chemiluminescent detection Fluorescent detection FIGURE 1 Principle of planar and bead-based multiplex detection and quantitation of cytokines, chemokines, growth factors, and other proteins. Assays use antibodies against (a-) various cytokines, and biotin (yellow)-streptavidin (green) conjugation. See text for details. [Adapted from H. Skalnikova et al., Mapping of the secretome of primary isolates of mammalian cells, stem cells and derived cell lines, 2011, Proteomics 11:691.] diction of nonclassically secreted proteins. Several bioinformatics tools, including TargetP and Protein Prowler, use the protein sequence to predict its subcellular localization. Finally, Ingenuity Pathway Analysis allows the investigator to search for protein interaction partners and to predict the involvement of the protein of interest in functional networks. Botto, S., D. N. Streblow, V. DeFilippis, L. White, C. N. Kreklywich, P. P. Smith, and P. Caposio. (2011). IL-6 in human cytomegalovirus secretome promotes angiogenesis and survival of endothelial cells through the stimulation of survivin. Blood 117:352–361. Receptors and Signaling: Cytokines and Chemokines TABLE 4-2 | CHAPTER 4 113 Six Cytokine Families Family name Representative members of family Comments Interleukin 1 family IL-1, IL-1, IL-1Ra, IL-18, IL-33 IL-1 was the first noninterferon cytokine to be identified. Members of this family include important inflammatory mediators. Hematopoietin (Class I cytokine) family IL-2, IL-3, IL-4, IL-5, IL-6, IL-7, IL-12, IL-13, IL15, IL-21, IL-23, GM–CSF, G-CSF, Growth hormone, Prolactin, Erythropoietin/hematopoietin This large family of small cytokine molecules exhibits striking sequence and functional diversity. Interferon (Class II cytokine) family IFN-, IFN-, IFN-, IL-10, IL-19, IL-20, IL-22, IL-24 While the IFNs have important roles in anti-viral responses, all are important modulators of immune responses. Tumor Necrosis Factor family TNF-, TNF-, CD40L, Fas (CD95), BAFF, APRIL, LT Members of this family may be either soluble or membrane bound; they are involved in immune system development, effector functions, and homeostasis. Interleukin 17 family IL-17 (IL17-A), IL17B, C, D, and F This is the most recently discovered family; members function to promote neutrophil accumulation and activation, and are proinflammatory. Chemokines (see Appendix III) IL-8, CCL19, CCL21, RANTES, CCL2 (MCP-1), CCL3 (MIP-1) All serve chemoattractant function. some general principles of cytokine and receptor architecture and function that should then enable the reader to place any cytokine into its unique biological context. Detailed studies of cytokine structure and function have revealed common features among families of cytokines. Cytokines are relatively small proteins and generally have a molecular mass of less than 30 kDa. Many are glycosylated, and glycosylation appears to contribute to cytokine stability, although not necessarily to cytokine activity. Cytokines characterized so far belong to one of six groups: the Interleukin 1 (IL-1) family, the Hematopoietin (Class I cytokine) family, the Interferon (Class II cytokine) family, the Tumor Necrosis Factor (TNF) family, the Interleukin 17 (IL-17) family, and the Chemokine family (Table 4-2). Each of these six families of cytokines, the receptors that engage them, and the signaling pathways that transduce the message received upon cytokine binding into the appropriate biological outcome are described in the following pages. Cytokines of the IL-1 Family Promote Proinflammatory Signals Cytokines of the interleukin 1 (IL-1) family are typically secreted very early in the immune response by dendritic cells and monocytes or macrophages. IL-1 secretion is stimulated by recognition of viral, parasitic, or bacterial antigens by innate immune receptors. IL-1 family members are generally proinflammatory, meaning that they induce an increase in the capillary permeability at the site of cytokine secretion, along with an amplification of the level of leukocyte migration into the infected tissues. In addition, IL-1 has systemic (whole body) effects and signals the liver to produce acute phase proteins such as the Type I interferons (IFNs  and ), IL-6, and the chemokine CXCL8. These proteins further induce multiple protective effects, including the destruction of viral RNA and the generation of a systemic fever response (which helps to eliminate many temperature-sensitive bacterial strains). IL-1 also activates both T and B cells at the induction of the adaptive immune response. Cytokines of the IL-1 Family Members of the IL-1 cytokine and receptor family are shown in Figure 4-5. The canonical (most representative) members of the IL-1 family, IL-1 and IL-1, are both synthesized as 31 kDa precursors, pro-IL-1 and pro-IL-1. Pro IL-1 is biologically active, and often occurs in a membranebound form, whereas pro-IL-1 requires processing to the fully mature soluble molecule before it can function. Pro-IL-1 and  are both trimmed to their 17 kDa active forms by the proteolytic enzyme caspase-1 inside the secreting cell. Active caspase-1 is located in a complex set of proteins referred to as the inflammasome (see Chapter 5). Other IL-1 family members, IL-18 and IL-33, have also been shown to be processed by caspase-1 in vitro (although there is ambiguity as to whether IL-33 requires this processing for full activity in vivo). IL-18 is related to IL-1, uses the same receptor family, and has a similar function; like IL-1, IL-18 is expressed by monocytes, macrophages, and dendritic cells and is secreted early in the immune response. In contrast, IL-33 is constitutively expressed in smooth muscle and in bronchial epithelia, and its expression can be induced by IL-1 and TNF- in lung and skin fibroblasts. IL-33 has been shown to induce TH2 cytokines that promote T-lymphocyte interactions with B cells, mast cells, and eosinophils. IL-33 PA R T I 114 | Introduction has also been implicated in the pathology of diseases such as asthma and inflammatory airway and bowel diseases. Two additional members of this cytokine family act as natural inhibitors of IL-1 family function. The soluble protein IL-1Ra (IL-1 Receptor antagonist) binds to the IL-1RI receptor, but prevents its interaction with its partner receptor chain, IL-1RAcP, thus rendering it incapable of transducing a signal to the interior of the cell. IL-1Ra therefore functions as an antagonist ligand of IL-1. IL-18BP adopts a different strategy of inhibition, binding to IL-18 in solution and preventing IL-18 from interacting productively with its receptor. The inhibitory effect of IL-18BP is enhanced by the further binding of IL-1F7 (see Figure 4-5b). The Interleukin 1 family of receptors includes the receptors for IL-1, IL-18, and IL-33. Both forms of IL-1—IL-1 and IL-1—bind to the same receptors and mediate the same responses. Two different receptors for IL-1 are known, and both are members of the immunoglobulin superfamily of proteins (see Chapter 3). Only the type I IL-1R (IL-1RI), which is expressed on many cell types, is able to transduce a cellular signal; the type II IL-1R (IL-1RII) is limited to B cells and is inactive. For full functioning, the Type 1 IL-1R also requires the presence of an interacting accessory protein, IL1RAcP (IL-1 Receptor Accessory Protein (see Figure 4-5a). (a) (b) The IL-1 Family of Cytokine Receptors IL-18 Inhibitory ligand and receptors IL-1Ra IL-1 Inhibitory ligands S S IL-1RI S S S S S S S S S S S S S S S S IL-1RAcP S S S S S S IL-18Rα S S sIL-1RII S S S S S S S S S S S S S S IL-18Rβ IL-18 sIL-1RAcP Blocked binding S S Cytoplasm Cytoplasm TIR domains IL-18BP IL-1F7 TIR domains IL-1RII Signal Signal (c) IL-33 Inhibitory receptors T1/ST2 S S S S S S S S S S S S Cytoplasm IL-1RAcP TIR domains Signal S S S S S S S S S S S S sST2 sIL-1RAcP FIGURE 4-5 Ligands and receptors of the IL-1 family. (a) The two agonist ligands, IL-1 and IL-1, are represented by IL-1 and the antagonist ligand by IL-1Ra. The IL-1 receptor, IL-1RI, has a long cytoplasmic domain and, along with IL-1RAcP, activates signal transduction pathways. IL-1Ra functions as an IL-1 inhibitor by binding to IL-1RI while not allowing interaction with IL-1RAcP. IL-1RII does not activate cells but functions as an IL-1 inhibitor both on the plasma membrane and in the cell microenvironment as a soluble receptor (sIL-1RII). IL-1RAcP can also inhibit IL-1 signals by cooperating with IL-1RII in binding IL-1 either on the plasma membrane or as a soluble molecule (sIL-1RAcP). (b) IL-18 binds to the IL-18R chain, and this complex then engages the IL-18R chain to initiate intracellular signals. The soluble protein IL-18BP functions as an inhibitor of IL-18 by binding this ligand in the fluid phase, preventing interaction with the IL-18R chain. IL-1F7 appears to enhance the inhibitory effect of IL-18BP. (c) IL-33 binds to the T1/ST2 receptor, and this complex engages the IL-1RAcP as a co-receptor. A soluble form of ST2 (sST2) may function as an inhibitor of IL-33 by binding IL-33 in the cell microenvironment and sIL-1RAcP may enhance the inhibitory effects of sST2. | Receptors and Signaling: Cytokines and Chemokines Note that both the IL-1RI and the IL-1RII receptor chains as well as the receptor accessory protein exist in both soluble and membrane-bound forms. However, a full signal is transmitted only from the dimer of the membrane-bound forms of IL-1RI and IL-1RAcP. The alternative, membrane-bound and soluble forms of IL-1 binding proteins, can “soak up” excess cytokine, but they are unable to transduce the interleukin signal. Thus, by secreting more or fewer of these inactive receptors, at different times during an immune response, the organism has the opportunity to fine-tune the cytokine signal by allowing the inactive and soluble receptors to compete with the signal-transducing receptor for available cytokine. This theme finds echoes in other immune system receptor families, and appears to be a frequently evolved strategy for controlling the strength of signals that give rise to important outcomes. In the case of IL-1, the ultimate result of successful IL-1 signaling is a global, proinflammatory state, and so the penalty paid by the host for an inappropriately strong IL-1 response would be physiologically significant and even potentially fatal. The receptor for IL-18 is also a heterodimer, made up of IL-18R and IL-18R. IL-33 is recognized by the IL-1RAcP in combination with a novel receptor protein, variously termed T1/ST-2 or IL-1RL1. As for IL-1, inhibitory receptors exist for IL-33 (see Figure 4-5c). Signaling from IL-1 Receptors Productive ligand binding to the extracellular portion of the IL-1 receptor leads to a conformational alteration in its cytoplasmic domain. This structural alteration in the receptor leads to a series of downstream signaling events (Figure 4-6). Most of the themes of these events will be familiar to the reader from Chapter 3, and we will encounter them again in the discussion of innate immune receptors in Chapter 5. First, binding of the adapter protein MyD88 to the occupied receptor allows recruitment to the receptor complex of one or more members of the IL-1 Receptor Activated Kinase (IRAK) protein family. One of these, IRAK-4, is activated by autophosphorylation and phosphorylates its fellow IRAKs, resulting in the generation of binding sites for TNF Receptor Associated Factor 6 (TRAF6), which is associated with a ubiquitin-ligase complex capable of generating polyubiquitin chains. The IRAK-TRAF6 complex now dissociates from the receptor and interacts with a preformed cytosolic complex made up of the kinase TGFAssociated Kinase 1 (TAK1) and two TAK1-Binding proteins, TAB1 and TAB2. Binding of polyubiquitin chains to the TAB proteins in the TAK1 complex activates it. The TAK1 complex now performs two functions with which the reader should be familiar. It phosphorylates and activates the IKK complex, leading to the destruction of IB and the resultant activation of the transcription factor NF-B (see Figure 3-17). In addition, TRAF6 also plays a role in IKK activation by providing ubiquitination sites to which the NEMO component of IKK can bind, resulting in its further activation. TAK1 also activates downstream members CHAPTER 4 115 IL-1 IL-1RI S S S S S S S S S S S S IL-1RAcP Cytoplasm TIR domains MyD88 IRAKs TRAF6 TAB1 Ub TAB2 TAK1 MAP kinase cascade NF-κB pathway AP-1 NF-κB Cytokine expression FIGURE 4-6 Signaling from members of the IL-1 receptor family. IL-1 binding to its receptor induces a conformational alteration in the receptor’s Toll-IL-1R (TIR) domain that allows binding of the adapter protein MyD88 via its TIR domain. MyD88 recruits one or more IL-1 receptor activated kinases (IRAKs) to the receptor complex, which phosphorylate one another providing binding sites for TRAF6. The IRAK-TRAF6 complex dissociates from the receptor complex and interacts with the cytoplasmic protein TAK1 and its two binding proteins, TABs 1 and 2. TRAF6, together with a ubiquitin-ligase complex, catalyzes the generation of polyubiquitin chains that activate the TAK1 complex. TAK1 activates downstream events leading to the activation and nuclear localization of the transcription factor NF-B. TAK1 also activates downstream members of the MAP kinase cascade, leading to activation of the AP-1 transcription factor. of the MAP kinase cascade, which in turn activate the AP-1 transcription factor (see Figure 3-16). Binding of IL-1 family cytokines to their receptors thereby leads to a global alteration in the transcription patterns of the affected cells, which in turn results in the up-regulation of proinflammatory cytokines and adhesion molecules. 116 PA R T I | Introduction Hematopoietin (Class I) Family Cytokines Share Three-Dimensional Structural Motifs, but Induce a Diversity of Functions in Target Cells Members of the hematopoietin (Class I) cytokine family are small, soluble cytokines that communicate between and among cells of the immune system. Their name is somewhat misleading in that not all members of this family are implicated in hematopoietic (blood-cell forming) functions per se. However, some of the earliest members of this family to be characterized indeed have hematopoietic functions, and the cytokine family was then defined on the basis of structural similarities among all the participants. Because the hematopoietin family contains some of the earliest cytokines to be structurally characterized, it is sometimes also referred to as the Class I cytokine family. Cytokines of the Hematopoietin (Class I) Family As more hematopoietin family members have been defined, it has become clear that their cellular origins and target cells are as diverse as their ultimate functions, which range from signaling the onset of T- and B-cell proliferation (e.g., IL-2), to signaling the onset of B-cell differentiation to plasma cells and antibody secretion (e.g., IL-6), to signaling the differentiation of a T helper cell along one particular differentiation pathway versus another (e.g., IL-4 vs. IL-12) and, finally, to initiating the differentiation of particular leukocyte lineages (e.g., GM-CSF, G-CSF). Appendix II lists the cytokines described in this book, along with their cells of origin, their target cells, and the functions they induce. Significant homology in the three-dimensional structure of hematopoietin family cytokines defines them as members of a single protein family, despite a relatively high degree of amino acid sequence diversity. The defining structural feature of this class of cytokines is a four-helix bundle motif, organized into four anti-parallel helices (Figure 4-7). Members of this family can then be further subclassified on the basis of helical length. Cytokines such as IL-2, IL-4, and IL-3 typically have short helices of 8 to 10 residues in length. In contrast, the so-called long-chain cytokines, which include IL-6 and IL-12, typically have helical lengths of 10 to 20 residues. The Hematopoietin or Class I Receptor Family Most hematopoietin cytokine receptors include two types of protein domains: an immunoglobulin-like domain, made up of  sheets, as described in Chapter 3, and domains that bear structural homology to the FNIII domain of the extracellular matrix protein fibronectin. Binding sites for most cytokines are to be found in a structure made up of two, tandem (side-by-side) FNIII domains referred to as Cytokine-binding Homology Regions (CHRs). As we will see, the CHR motif is common to cytokine receptors from several families. A feature common to most of the Hematopoietin and Interferon cytokine receptor families is the presence of multiple subunits. Table 4-3 lists the three subfamilies of hematopoietin (a) 36 114 30 α 42 73 80 58 D 133 C A 6 C 96 B S S 105 52 N (b) FIGURE 4-7 The four-helix bundle is the defining structural feature of the Hematopoietin family of cytokines. Structure of interleukin 2—the defining member of the Hematopoietin family—showing the four -helices of the hematopoietin cytokines point in alternating directions. (a) Topographical representation of the primary structure of IL-2 showing -helical regions ( and A-D) and connecting chains of the molecule. (b) Ribbon representation of the crystallographic structure of human Il-2. [Part (b) PDB ID 1M47.] receptors, each subfamily being defined by a receptor subunit that is shared among all members of that family. The -Chain Bearing, or IL-2 Receptor, Subfamily Expression of a common  chain defines the IL-2 receptor subfamily, which includes receptors for IL-2, IL-4, IL-7, IL-9, IL-15, and IL-21. The IL-2 and the IL-15 receptors are heterotrimers, consisting of a cytokine-specific  chain and two chains— and —responsible for both signal transduction and cytokine recognition. The IL-2 receptor  chain also functions as the signal-transducing subunit for the other receptors in this subfamily, which are all dimers. Congenital X-linked severe combined immunodeficiency (XSCID) results from a defect in the -chain gene, which maps to the X chromosome. The immunodeficiency observed in this disorder, which includes deficiencies in both T-cell and | Receptors and Signaling: Cytokines and Chemokines TABLE 4-3 Subfamilies of hematopoietin family cytokine receptors share common subunits Common cytokine receptor subunit Cytokines recognized by receptors bearing that common subunit  IL-2, IL-4, IL-7, IL-9, IL-15, IL-12  IL-3, IL-5, GM-CSF gp130 IL-6, IL-11, LIF, OSM, CNTF, IL-27 (a) Intermediate affinity, IL-2R High affinity, IL-2R Dissociation constant (Kd): Cells expressed by: (b) α β IL-2Rβ IL-2Rγ IL-2Rα IL-2Rβ IL-2Rγ IL-2Rα 10−9 M 10−11 M 10−8 M NK cells Resting T cells (low numbers) Activated CD4+ and CD8+ T cells Activated B cells (low numbers) IL-2Rα IL-2 Low affinity, IL-2R γ β Subunit composition: 117 NK-cell activity, results from the loss of all the cytokine functions mediated by the IL-2 subfamily receptors. The IL-2 receptor occurs in three forms, each exhibiting a different affinity for IL-2: the low-affinity monomeric IL-2R (CD25) (which can bind to IL-2, but is incapable of transducing a signal from it), the intermediate-affinity dimeric IL-2R (which is capable of signal transduction), and the high-affinity trimeric IL-2R (which is responsible for most physiologically relevant IL-2 signaling) (Figure 4-8a). A recent x-ray crystallographic structure of the high-affinity trimeric form of the IL-2 receptor with an IL-2 molecule in its binding site reveals that IL-2 binds in a pocket formed by the  and  chains α γ CHAPTER 4 IL-2Rα γc IL-2 IL-2Rβ IL-2Rβ FIGURE 4-8 Comparison of the three forms of the IL-2 receptor. (a) Schematic of the three forms of the receptor and listing of dissociation constants and properties for each. Signal transduction is mediated by the  and  chains, but all three chains are required for high-affinity binding of IL-2. (b) Three-dimensional structure of the γc three-chain form of the IL-2 receptor with bound IL-2 (views rotated by 90). Note that the  chain completes the pocket to which IL-2 binds, accounting for the higher affinity of the trimeric form. [From X. Wang, M. Rickert, and K. C. Garcia, 2005, Structure of the quaternary complex of interleukin–2 with its , , and c receptors. Science 310:1159.] 118 PA R T I | Introduction (Figure 4-8b). Important additional contacts with the IL-2 ligand are contributed when the  chain is present, accounting for the higher affinity of binding by the trimer. The expression of the three chains of the IL-2 receptor varies among cell types and in different activation states. The intermediate affinity () IL-2 receptors are expressed on resting T cells and on NK cells, whereas activated T and B cells express both the low-affinity () and the high-affinity () receptor forms (see Figure 4-8a). Since there are approximately ten times as many low-affinity as high-affinity receptors on activated T cells (50,000 vs. 5000), one must ask what the function of the low-affinity receptor might be, and two possible ideas have been advanced. It may serve to concentrate IL-2 onto the recipient cell surface for passage to the high-affinity receptor. Conversely, it may reduce the local concentration of available IL-2, ensuring that only cells bearing the high-affinity receptor are capable of being activated. Whatever the answer to this question may be, the restriction of the high-affinity IL-2 receptor expression to activated T cells ensures that only antigen-activated CD4 and CD8 T cells will proliferate in response to physiologic levels of IL-2. The Chain Bearing, or GM-CSF, Receptor Subfamily Members of the GM-CSF receptor subfamily, which includes the receptors for IL-3, IL-5, and GM-CSF, share the  signaling subunit. Each of these cytokines binds with relatively low affinity to a unique, cytokine-specific receptor protein, the  subunit of a dimeric receptor. All three low-affinity subunits associate noncovalently with the common signal-transducing  subunit. The resulting  dimeric receptor has a higher affinity for the cytokine than the specific  chain alone, and is also capable of transducing a signal across the membrane upon cytokine binding (Figure 4-9a). IL-3, IL-5, and GM-CSF exhibit redundant activities. IL-3 and GM-CSF both act on hematopoietic stem cells and progenitor cells, activate monocytes, and induce megakaryocyte differentiation, and all three of these cytokines induce eosinophil proliferation and basophil degranulation with release of histamine. Since the receptors for IL-3, IL-5, and GM-CSF share a common signal-transducing  subunit, each of these cytokines would be expected to transduce a similar activation signal, accounting for the redundancy seen among their biological effects, and indeed, all three cytokines induce the same patterns of protein phosphorylation upon cell activation. However, when introduced simultaneously to a cell culture, IL-3 and GMCSF appear to antagonize one another; the binding of IL-3 is inhibited by GM-CSF, and binding of GM-CSF is inhibited by IL-3. This antagonism is caused by competition for a limited number of  subunits available to associate with the cytokinespecific  subunits of the dimeric receptors (Figure 4-9b). The gp130 Receptor Subfamily The importance of the gp130 cytokine receptor family to the development and health of the individual is underscored by the results of deletion studies which have demonstrated that (a) IL-3 IL-5 GM-CSF Low-affinity receptors Exterior Membrane α α α β subunit α α α Interior High-affinity receptors β (b) β β IL-3 α α GM-CSF α α α α β α α β α β β β α α α β FIGURE 4-9 Interactions between cytokine-specific subunits and a common signal-transducing subunit of the -chain family of cytokine receptors. (a) Schematic diagram of the low-affinity and high-affinity receptors for IL-3, IL-5, and GM-CSF. The cytokine-specific subunits exhibit low-affinity binding and cannot transduce an activation signal. Noncovalent association of each subunit with a common  subunit yields a high-affinity dimeric receptor that can transduce a signal across the membrane. (b) Competition of ligand-binding chains of different receptors for a common subunit can produce antagonistic effects between cytokines. Here binding of IL-3 by  subunits of the IL-3 receptor allows them to outcompete  chains of the GM-CSF receptor for  subunits. [Part a adapted from T. Kishimoto et al., 1992, Interleukin-6 and its receptor: A paradigm for cytokines, Science 258:593.] the targeted disruption of gp130 in mice during embryonic development is lethal. Receptors in this family include those for IL-6, important in the initiation of the immune response, and IL-12, critical for signaling differentiation of helper T cells along the TH1 pathway. Targeted disruption of individual cytokine receptors, as well as of the cytokines themselves, has provided much functional information about signaling via these cytokine family members. Cytokine specificity of the gp130 family of receptors is determined by the regulated expression of ligand-specific chains in dimers, or trimers, with the gp130 component. The Receptors and Signaling: Cytokines and Chemokines gp130 subunit family of cytokine receptors is further subdivided into receptors specific for monomeric cytokines, such as IL-6, and those which bind the dimeric cytokines, such as IL-12. Because the Hematopoietin (Class I) and Interferon (Class II) cytokine receptor families utilize similar signaling pathways, we will first describe the Interferons and their receptors and then consider the signaling pathways used by the two families together. The Interferon (Class II) Cytokine Family Was the First to Be Discovered In the late 1950s, investigators studying two different viral systems in two laboratories half a world apart almost simultaneously discovered interferons. Yasu-Ichi Nagano and Yashuhiko Kojima, Japanese virologists, were using a rabbit skin and testes tissue culture model to develop a vaccine against smallpox. They noted that immunization with a UV-inactivated form of the cowpox virus resulted in the localized inhibition of viral growth, following a subsequent injection of the same virus. Viral growth inhibition was restricted to a small area of skin close to the site of the original immunization, and the scientists postulated that the initial injection had resulted in the production of a “viral inhibitory factor.” After showing that their “inhibitory factor” was not simply antibody, they published a series of papers about it. With hindsight, scientists now believe that their protective effect was mediated by interferons. However, the technical complexity of their system, and the fact that their papers were published in French, rather than in English, delayed the dissemination of their findings to the broader scientific community. Meanwhile, in London, Alick Isaacs and Jean Lindenman were growing live influenza virus on chick egg chorioallantoic membranes (a method that is still used today), and noticed that exposure of their membranes to a heat-inactivated form of influenza interfered with subsequent growth of a live virus preparation on that surface preparation. They proved that the growth inhibition resulted from the production of an inhibiting molecule by the chick membrane. They named it “interferon” because of its ability to “interfere” with the growth of the live virus. Their more straightforward in vitro assay system enabled them to rapidly characterize the biological effects of the molecule involved, and they wrote a series of papers describing the biological effects of interferon(s) in the late 1950s. However, since interferons are active at very low concentrations, it was not until 1978 that they were produced in quantities sufficient for biochemical and crystallographic analysis. Since that time, investigators have shown that there are two major types of interferons, Types 1 and 2, and that Type 1 interferons can be subdivided into two subgroups. Interferons Type I interferons are composed of Interferons , a family of about 20 related proteins, and interferon-, which are secreted by activated macrophages and dendritic cells, as well as by virus-infected cells. Interferons  and  are also secreted by | CHAPTER 4 119 virally infected cells after recognition of viral components by pattern recognition receptors (PRRs) located either at the cell surface, or inside the cell (see Chapter 5). Intracellular PRRs may interact with virally derived nucleic acids or with endocytosed viral particles. The secreted Type I interferons then interact in turn with membrane-bound interferon receptors on the surfaces of many different cell types. The results of their interaction with these receptors are discussed in detail in Chapter 5, but they include the induction of ribonucleases that destroy viral (and cellular) RNA, and the cessation of cellular protein synthesis. Thus, interferons prevent virally infected cells from replicating and from making new viral particles. However, they simultaneously inhibit normal cellular functions and destroy virally infected cells so that the infection cannot spread. Type I interferons are dimers of 18 to 20 kDa polypeptides, predominantly helical in structure, and some members of this family are naturally glycosylated. Type I interferons are used in the treatment of a variety of human diseases, most notably hepatitis infections. Type II interferon, otherwise known as interferon-, is produced by activated T and NK cells and is released as a dimer (Figure 4-10). Interferon- is a powerful modulator of FIGURE 4-10 The complex between IFN- and the ligandbinding chains of its receptor. This model is based on the x-ray crystallographic analysis of a crystalline complex of interferon- (dark and light purple) bound to ligand-binding  chains of the receptors (green and yellow). Note that IFN- is shown in its native dimeric form; each member of the dimer engages the  chain of an IFN- receptor, thereby bringing about receptor dimerization and signal transduction. [From M. R. Walter et al., 1995, Crystal structure of a complex between interferon- and its soluble high-affinity receptor. Nature 376:230, courtesy M. Walter, University of Alabama.] 120 PA R T I | Introduction CLINICAL FOCUS Therapy with Interferons Interferons are an extraordinary group of proteins with important effects on the immune system. Their actions affect both the adaptive and the innate arms of the immune system and include the induction of increases in the expression of both Class I and Class II MHC molecules and the augmentation of NK-cell activity. Cloning of the genes that encode IFN-, IFN-, and IFN- has made it possible for the biotechnology industry to produce large amounts of each of these interferons at costs that make their clinical use practical (Table 1). IFN- (also known by its trade names Roferon and Intron A) has been used for the treatment of hepatitis C and hepatitis B. It has also been found useful in a number of different applications in cancer therapy. A type of B-cell leukemia known as hairy-cell leukemia (because the cells are covered with fine, hairlike cytoplasmic projections) responds well to IFN-. Chronic myelogenous leukemia, a disease characterized by increased numbers of granulocytes, begins with a slowly developing chronic phase that changes to an accelerated phase and terminates in a blast phase, which is usually resistant to treatment. IFN- is an effective treatment for this leukemia in the chronic phase (70% response rates have been reported), and some patients (as many as 20% in some studies) undergo complete remission. Kaposi’s sarcoma, the cancer most often seen in AIDS patients in the United States, also responds to treatment with IFN-, and there are reports of a trend toward longer survival and fewer oppor- tunistic infections in patients treated with this agent. Most of the effects mentioned above have been obtained in clinical studies that used IFN- alone, but certain applications such as hepatitis C therapy commonly use it with an antiviral drug such as ribavirin. The clearance time of IFN- is lengthened by using it in a form complexed with polyethylene glycol (PEG) called pegylated interferon. IFN- has emerged as the first drug capable of producing clinical improvement in multiple sclerosis (MS). Young adults are the primary target of this autoimmune neurologic disease, in which nerves in the central nervous system (CNS) undergo demyelination. This results in progressive neurologic dysfunction, leading to significant and, in many cases, severe disability. This disease is often characterized by periods of nonprogression and remission alternating with periods of relapse. Treatment with IFN- provides longer periods of remission and reduces the severity of relapses. Furthermore, magnetic resonance imaging (MRI) studies of CNS damage in treated and untreated patients revealed that MSinduced damage was less severe in a group of IFN-treated patients than in untreated ones. IFN- has been used, with varying degrees of success, to treat a variety of malignancies that include non-Hodgkin’s lymphoma, cutaneous T-cell lymphoma, and multiple myeloma. A more successful clinical application of IFN- in the clinic is in the treatment of the hereditary immunodeficiency chronic granulomatous dis- the adaptive immune response, biasing T cell help toward the TH1 type and inducing the activation of macrophages, with subsequent destruction of any intracellular pathogens and the differentiation of cytotoxic T cells. All three interferons increase the expression of MHC complex proteins on the surface of cells, thus enhancing their antigen-presentation capabilities. ease (CGD; see Chapter 18). CGD features a serious impairment of the ability of phagocytic cells to kill ingested microbes, and patients with CGD suffer recurring infections by a number of bacteria (Staphylococcus aureus, Klebsiella, Pseudomonas, and others) and fungi such as Aspergillus and Candida. Before interferon therapy, standard treatment for the disease included attempts to avoid infection, aggressive administration of antibiotics, and surgical drainage of abscesses. A failure to generate microbicidal oxidants (H2O2, superoxide, and others) is the basis of CGD, and the administration of IFN- significantly reverses this defect. Therapy of CGD patients with IFN- significantly reduces the incidence of infections. Also, the infections that are contracted are less severe, and the average number of days spent by patients in the hospital is reduced. IFN- has also been shown to be effective in the treatment of osteopetrosis (not osteoporosis), a life-threatening congenital disorder characterized by overgrowth of bone that results in blindness and deafness. Another problem presented by this disease is that the buildup of bone reduces the amount of space available for bone marrow, and the decrease in hematopoiesis results in fewer red blood cells and anemia. The decreased generation of white blood cells causes an increased susceptibility to infection. The use of interferons in clinical practice is likely to expand as more is learned about their effects in combination with other therapeutic agents. Interferon- is used medically to bias the adaptive immune system toward a cytotoxic response in diseases such as leprosy and toxoplasmosis, in which antibody responses are less effective than those that destroy infected cells. Clinical Focus Box 4-2 describes additional roles of interferons in the clinic. Minority members of the Interferon family of cytokines include IL-10, secreted by monocytes and by T, B, and Receptors and Signaling: Cytokines and Chemokines | CHAPTER 4 121 BOX 4-2 TABLE 1 Cytokine-based therapies in clinical use Agent Nature of agent Clinical application Enbrel Chimeric TNF-receptor/IgG constant region Rheumatoid arthritis Remicade or Humira Monoclonal antibody against TNF- receptor Rheumatoid arthritis, Crohn’s disease Roferon Interferon--2a* Hepatitis B, Hairy-cell leukemia, Kaposi’s sarcoma, Hepatitis C† Intron A Interferon-–2b Melanoma Betaseron Interferon-–1b Multiple sclerosis Avonex Interferon-–1a Multiple sclerosis Actimmune Interferon-–1b Chronic granulomatous disease (CGD), Osteopetrosis Neupogen G-CSF (hematopoietic cytokine) Stimulates production of neutrophils; reduction of infection in cancer patients treated with chemotherapy, AIDS patients Leukine GM-CSF (hematopoietic cytokine) Stimulates production of myeloid cells after bone marrow transplantation Neumega or Neulasta Interleukin 11 (IL-11), a hematopoietic cytokine Stimulates production of platelets Epogen Erythropoietin (hematopoietic cytokine) Stimulates red-blood-cell production Ankinra (kineret) Recombinant IL-1Ra Rheumatoid arthritis Daclizumab (Zenapax) Humanized monoclonal antibody against IL-2R Prevents rejection after transplantation Basiliximab (Simulect) Human/mouse chimeric monoclonal antibody against IL-2R Prevents transplant rejection *Interferon-–2a is also licensed for veterinary use to combat feline leukemia. † Normally used in combination with an antiviral drug (ribavirin) for hepatitis C treatment. Although interferons, in common with other cytokines, are powerful modifiers of biological responses, the side effects accompanying their use are fortunately relatively mild. Typical side effects include flu-like symptoms, such as headache, fever, chills, and fatigue. These symptoms can largely be managed with acetaminophen (Tylenol) and diminish in intensity during continued treatment. Although dendritic cells that regulates immune responses. IL-10 shares structural similarities with interferon-, and these similarities enable it to bind to the same class of receptors. In addition, a third class of interferons, the so-called interferon- , or type III Interferon family, was discovered in 2003. There are currently three members of this family: interferon- 1 (IL-29), interferon- 2 (IL-28A), and interferon- 3 (IL-28B). interferon toxicity is usually not severe, treatment is sometimes associated with serious manifestations such as anemia and depressed platelet and white-bloodcell counts. Like Type I interferons, the Type III interferons up-regulate the expression of genes controlling viral replication and host cell proliferation. Interferon Receptors Members of the Interferon receptor family are heterodimers that share similarly located, conserved cysteine residues with 122 PA R T I | Introduction members of the Hematopoietin receptor family. Initially, only interferon-, -, and - were thought to be ligands for these receptors. However, recent work has shown that the receptor family consists of 12 receptor chains that, in their various assortments, bind no fewer than 27 different cytokines, including six members of the IL-10 family, 17 Type I interferons, one Type II interferon, and three members of the recently described interferon– family, including IL-28A, IL-28B, and IL-29. The JAK-STAT Signaling Pathway Early experiments in cytokine signaling demonstrated that a series of protein tyrosine phosphorylations rapidly followed the interaction of a cytokine with a receptor from the Class I or Class II cytokine receptor families. These results were initially puzzling, since the cytokine receptors lack the immunotyrosine activation motifs (ITAMs) characteristic of B- and T-cell receptors. However, studies of the molecular events triggered by binding of interferon gamma (IFN-) to its receptor shed light on the mode of signal transduction used by members of both the Hematopoietin and Interferon cytokine families. In the absence of cytokine, the receptor subunits are associated only loosely with one another in the plane of the membrane, and the cytoplasmic region of each of the receptor subunits is associated noncovalently with inactive tyrosine kinases named Janus Activated Kinases (JAKs). (Some members of this family of kinases retain their earlier name of Tyk, but share structural and functional properties with the JAK family of kinases.) The process of signal transduction from Class I and Class II cytokine receptors has been shown to proceed according to the following steps (Figure 4-11): • Cytokine binding induces the association of the two separate cytokine receptor subunits and activation of the receptor-associated JAKs. • The receptor-associated JAKs phosphorylate specific tyrosines in the receptor subunits. • These phosphorylated tyrosine residues serve as docking sites for inactive transcription factors known as Signal Transducers and Activators of Transcription (STATs). • The inactive STATs are phosphorylated by JAK and Tyk kinases. • Phosphorylated STAT transcription factors dimerize, binding to one another via SH2/phosphotyrosine interactions. • Phosphorylation also results in a conformational change in the STAT dimer that reveals a nuclear localization signal. • The STAT dimer translocates into the nucleus, where it initiates the transcription of specific genes. Currently, we know of seven STAT proteins (STAT 1-4, 5A, 5B, and 6) and four JAK proteins (JAK 1–3 and Tyk2) in Cytokine α β β α Dimerization of receptor P JAK P P SH2 P P Activation of JAK family tyrosine kinases, phosphorylation of receptor P Tyrosine phosphorylation of STAT by JAK kinase STAT P P Dimerization of STAT P P DNA Specific gene transcription FIGURE 4-11 General model of signal transduction mediated by most Class I and Class II cytokine receptors. Binding of a cytokine induces dimerization of the receptor subunits, which leads to the activation of receptor subunit-associated JAK tyrosine kinases by reciprocal phosphorylation. Subsequently, the activated JAKs phosphorylate various tyrosine residues, resulting in the creation of docking sites for STATs on the receptor and the activation of one or more STAT transcription factors. The phosphorylated STATs dimerize and translocate to the nucleus, where they activate transcription of specific genes. mammals. Specific STATs play essential roles in the signaling pathways of a wide variety of cytokines (Table 4-4). Given the generality of this pathway among Class I and Class II cytokines, how does the immune system induce a specific response to each cytokine? First, there is exquisite specificity in the binding of cytokines to their receptors. Secondly, particular cytokine receptors are bound to specific partner JAK enzymes that in turn activate unique STAT transcription factors. Third, the transcriptional activity of activated STATs is specific because a particular STAT homodimer or heterodimer will only recognize certain sequence motifs and thus can interact only with the promoters of certain genes. Finally, only those target genes whose expression is permitted by a particular cell type can be activated within that variety of cell. For example, promoter regions in some cell types may be caught up in heterochromatin and Receptors and Signaling: Cytokines and Chemokines TABLE 4-4 | CHAPTER 4 123 STAT and JAK interaction with selected cytokine receptors during signal transduction Each cytokine receptor must signal through a pair of Janus kinases. The JAKs may operate as either homo- or heterodimers. Cytokine receptor Janus kinase STAT IFN-/- JAK 1, Tyk 2* STATs 1 and 2 IFN- JAK 1, JAK 2 STAT 1 IL-2 JAK 1, JAK 3 Mainly STATs 3 and 5. Also STAT 1. IL-4 JAK 1, JAK 3 Mainly STAT 6. Also STAT 5. IL-6 JAK 1, JAK 2 STAT 3 IL-7 JAK 1, JAK 3 STATs 5 and 3 IL-12 JAK 2, Tyk2 STATS 2, 3, 4, and 5 IL-15 JAK 1, JAK 3 STAT 5 IL-21 JAK 1, JAK 3 Mainly STATs 1 and 3; also STAT 5 * Despite its name, Tyk2 is also a Janus kinase. inaccessible to transcription factors. In this way, the Class I cytokine IL-4 can induce one set of genes in T cells, another in B cells, and yet a third in eosinophils. JAK-STAT pathways are not unique to the immune system. Among the many genes known to be regulated by mammalian STAT proteins are those encoding cell survival factors such as the Bcl-2 family members, those involved in cell proliferation such as cyclin D1 and myc, and those implicated in angiogenesis or metastasis such as vascular endothelial growth factor, or VEGF. At the close of cytokine signaling, negative regulators of the STAT pathway, such as protein inhibitor of activated STAT (PIAS), suppressor of cytokine signaling (SOCS), and protein tyrosine phosphatases are believed to be responsible for turning off JAK-STAT signaling and returning the cell to a quiescent state. Members of the TNF Cytokine Family Can Signal Development, Activation, or Death The Tumor Necrosis Family (TNF) family of cytokines regulates the development, effector function, and homeostasis of cells participating in the skeletal, neuronal, and immune systems, among others. Cytokines of the TNF Family Can Be Soluble or Membrane Bound TNF-related cytokines are unusual in that they are often firmly anchored into the cell membrane. Generally they are Type 2 transmembrane proteins with a short, intracytoplasmic N-terminal region, and a longer, extracellular C-terminal region. The extracellular region typically contains a canonical TNF-homology domain responsible for interaction with the cytokine receptors. Members of the TNF family can also act as soluble mediators, following cleavage of their extracellular regions, and in some cases, the same cytokine exists in both soluble and membrane-bound forms. There are two eponymous (having the same name as) members of the TNF family: TNF- and TNF-, though TNF- is more commonly known as Lymphotoxin-, or LT-. Both of these are secreted as soluble proteins. TNF- (frequently referred to simply as TNF) is a proinflammatory cytokine, produced primarily by activated macrophages, but also by other cell types including lymphocytes, fibroblasts, and keratinocytes (skin cells), in response to infection, inflammation, and environmental stressors. TNF elicits its biological effects by binding to its receptors, TNF-R1 or TNF-R2, which are described below. Lymphotoxin- is produced by activated lymphocytes and can deliver a variety of signals. On binding to neutrophils, endothelial cells, and osteoclasts (bone cells), Lymphotoxin- delivers activation signals; in other cells, binding of Lymphotoxin- can lead to increased expression of MHC glycoproteins and of adhesion molecules. We will also encounter five physiologically significant, membrane-bound members of the TNF cytokine family throughout this book. Lymphotoxin-, a membrane-bound cytokine, is important in lymphocyte differentiation. We will learn about BAFF and APRIL in the context of B-cell development and homeostasis (Chapter 10). CD40L is a cytokine expressed on the surface of T cells that is required to signal for B-cell differentiation (Chapter 12). Fas ligand (FasL), or CD95L, induces apoptosis on binding to its cognate receptor, Fas, or CD95. Whether membrane-bound or in soluble form, active cytokines of the TNF family assemble into trimers. Although in most cases they are homotrimeric, heterotrimeric cytokines do form between the TNF family members Lymphotoxin- and Lymphotoxin- and between APRIL and BAFF. Crystallographic analysis of TNF family members has revealed that 124 PA R T I | Introduction tant questions still await resolution. One reason that these pathways have been so difficult to define is that the same receptor, TNF-R1, can transduce both activating and deathpromoting signals, depending on the local cellular and molecular environment in which the signal is received, and investigators have yet to determine the trigger that shifts the signaling program from life to death. However, much is known about how each of these signaling pathways work, once that all-important decision has been made. We will start by describing the proapoptotic (deathinducing) pathway that is initiated when the membranebound TNF family member FasL on one cell binds to a Fas receptor on a second cell, leading to death in the cell bearing the Fas receptor. With this as our foundation, we will then illustrate how the TNF-R1 receptor mediates both life- and death- promoting signals. Signaling through other TNF-R family members, such as CD40, BAFF, and April, will be described in later chapters in the context of the various immune responses in which they are involved. FIGURE 4-12 The TNF-family members act as trimers in vivo. [PDB ID 1TNF] they have a conserved tertiary structure and fold into a -sheet sandwich. The conserved residues direct the folding in the internal  strands that, in turn, promote the trimer formation (Figure 4-12). TNF Receptors Members of the TNF receptor superfamily are defined by the presence of Cysteine-Rich Domains (CRDs) in the extracellular, ligand-binding domain. Each CRD typically contains six cysteine residues, which form three disulfide bonded loops, and individual members of the superfamily can contain from one to six CRDs. Although most TNF receptors are Type 1 membrane proteins (their N-terminals are outside the cell), a few family members are cleaved from the membrane to form soluble receptor variants. Alternatively, some lack a membrane anchoring domain at all, or are linked to the membrane only by covalently bound, glycolipid anchors. These soluble forms of TNF family receptors are known as “decoy receptors,” as they are capable of intercepting the signal from the ligand before it can reach a cell, effectively blocking the signal. This is a theme that we have encountered before in our consideration of the IL-1 receptor family. Signaling Through TNF Superfamily Receptors The work of delineating the precise pathways of signaling through TNF family receptors is ongoing, and some impor- Signaling Through the Fas Receptor At the close of an immune response, when the pathogen is safely demolished and the immune system needs to eliminate the extra lymphocytes it has generated to deal with the invader, responding lymphocytes begin to express the TNF family receptor Fas on their cell surfaces. Fas, and its ligand FasL, are specialized members of the TNF receptor and the TNF cytokine families, respectively, and they work together to promote lymphocyte homeostasis. Mice with mutations in either the fas (mrl/lpr mice) or the fasL (gld mice) genes consequently suffer from severe lympho-proliferative disorders, indicative of their inability to eliminate lymphocytes that are no longer serving a useful purpose. On interaction with other immune cells bearing FasL, the Fas receptor trimerizes and transduces a signal to the interior of the Fas-bearing cell that results in its elimination by apoptosis. Apoptosis, or programmed cell death, is a mechanism of cell death in which the cell dies from within and is fragmented into membrane-bound vesicles that can be rapidly phagocytosed by neighboring macrophages (Figure 4-13a). By using such well-controlled apoptotic pathways, the organism ensures that minimal inflammation is associated with the natural end of an immune response. Activation of the apoptotic pathway invokes the activation of caspases; these are proteases, bearing Cysteine residues at their active sites, which cleave after ASPartic acid residues. Binding of Fas to FasL results in the clustering of the Fas receptors (Figure 4-13b). This, in turn, promotes interaction between their cytoplasmic regions, which include domains common to a number of proapoptotic signaling molecules called death domains. This type of interaction, between homologous protein domains expressing affinity for one another, is referred to as a homotypic interaction. As they bind to one another, the clustered Fas protein death domains incorporate death domains from the adapter Receptors and Signaling: Cytokines and Chemokines (a) Hoechst Transmission (b) | CHAPTER 4 125 (a) T cell FasL Control Death domains Fas Death effector domains Procaspase-8 (inactive) FADD Early Caspase-8 (active) Late Procaspase-3 and Procaspase-7 (inactive) Caspase-3 and Caspase-7 (active) Inactive pro-apoptotic enzymes Active apoptotic enzymes Apoptosis FIGURE 4-13 Apoptotic signaling through Fas receptors. domains also located on the FADD adapter proteins incorporate the DED domains of procaspase-8 into the membrane complex. Clustering of procaspase-8 induces cleavage of the pro domains of procaspase-8, leading to the release of the active caspase-8 protease. Caspase-8 cleaves the pro domains from the executioner caspases, caspase-3 and caspase-7, which in turn cleave and activate nucleases leading to the degradation of nuclear DNA. Caspase-8 also cleaves and activates the proapoptotic Bcl-2 family member protein, BID. [Taimen, Pekka and Kallajoki, Markku. (a) Human HeLa cells were activated to undergo apoptosis through the Fas receptors. Hoechst-stained cells show the gradual condensation of nuclear DNA into membrane-bounded blebs, as the cell breaks up into vesicular packages that are recognized and phagocytosed by macrophages in the absence of inflammation. The same cells are also shown under transmission microscopy. Arrows show staining of the Nuclear Mitotic Apparatus protein, an early nuclear caspase target in apoptosis. (b) Signaling from Fas leads to apoptosis. Binding of FasL to Fas induces clustering of the Fas receptors and corresponding clustering of the Fas Death Domains (DDs). The DDs of the adapter protein FADD bind to the clustered Fas DDs via a homotypic interaction. Death effector NuMA and nuclear lamins behave differently in Fasmediated apoptosis J Cell Sci 2003 116:(3):571-583; Advance Online Publication December 11, 2002, doi:10.1242/jcs.00227. Reproduced with permission of Journal of Cell S] protein FADD (Fas-Associated Death Domain-containing protein). FADD contains not only death domains, but also a related type of domain called a Death Effector Domain (DED). This, in turn, binds homotypically to the DED domains of procaspase-8, resulting in the clustering of procaspase-8 molecules. Procaspase-8 molecules contain the active caspase-8 enzyme, held in an inactive state by binding to prodomains. The multimerization of procaspase-8 molecules results in mutual cleavage of their prodomains and induces capsase-8 activation. Caspase-8 then cleaves many target proteins critical to the generation of apoptosis. The target proteins of caspase-8 include the executioner caspases, -3 and -7 (which cleave and activate nucleases leading to the degradation of nuclear DNA), and the proapoptotic Bcl-2 family member BID. The complex of Fas, FADD, and procaspase-8 is referred to as the Death-Inducing Signaling Complex (DISC). The ultimate result of activation of this cascade is the condensation of nuclear material (see Figure 4-13a), the degradation of nuclear DNA into 240 base pair, nucleic acid fragments, and the subsequent breakdown of the cell into “easily digestible” membrane-bound fragments. 126 PA R T I | Introduction Signaling Through the TNF-R1 Receptor The TNF-R1 receptor is present on the surface of all vertebrate cells and, like Fas, has an intracytoplasmic death domain (DD). Although this receptor is capable of binding to both TNF- and Lymphotoxin-, we will focus on the signaling that is elicited by TNF- (TNF). TNF binding to the TNF-R1 receptor can lead to two very different outcomes: apoptosis (death) or survival (life). How it does so is still the focus of intensive investigation, but the story as it is unfolding is already a fascinating one. The mechanism by which TNF binding leads to apoptosis is slightly different from that which follows Fas-FasL binding. Like FasL, binding of TNF to the TNF-R1 receptor induces trimerization of the receptor as well as an alteration in its conformation, and these together result in the binding of a DD-containing adapter molecule, in this case TRADD, to the internal face of the receptor molecule (Figure 4-14). The TRADD adapter molecule provides additional binding sites for the components RIP1 (a serine/threonine kinase, rather evocatively named Rest In Peace 1), which binds via its own DDs and TRAF2, the TNF Receptor Associated Factor 2. This is known as complex I. Intracellular localization experiments have shown that this complex can dissociate from the TNF-R at the membrane, and migrate to the cytoplasm where it binds to the now familiar DD-containing protein FADD. FADD recruits procaspase-8, as described above, resulting in the generation of an apoptotic signal. The proapoptotic cytoplasmic complex generated upon TNF-R1 receptor binding is shown in Figure 4-14a as complex II. Counterintuitively, binding of this same molecule, TNF, to the same receptor, TNF-R1, can result in the delivery of survival as well as of proapoptotic signals. How can the same cytokine, acting through the same receptor, bring about two apparently opposing actions? In the TNF-mediated survival pathway (Figure 4-14b), the generation of the original membrane complex appears to initiate in the same general manner as for the proapoptotic pathway. However, in the case of the prosurvival pathway, the TRADD-containing complex does not dissociate from the membrane, but rather remains at the cell surface and recruits a number of other components, including the ubiquitin ligases cIAP1 and cIAP2. Once cIAP1 and cIAP2 join the TNF-R1 complex at the cell membrane, they recruit the LUBAC proteins, which attach linear ubiquitin chains to RIP1. Polyubiquitinated RIP1 then binds to the NEMO component of IKK as well as to TAK1, which is already complexed with its associated TAB proteins as described above. RIP1 and TAK1 activate the IKK complex, leading to IB phosphorylation and destruction, and subsequent activation of NF-B. Once NF-B is fully activated, it turns on the expression of the cFLIP protein that then inhibits the activity of caspase-8. This effectively shuts down the antagonistic, proapoptotic pathway (Figure 4-14a). As previously described, the TAK1 complex also acts to activate the MAP kinase pathway, which further enhances survival signaling. (a) Apoptosis (b) Survival TNF TNF TNF-R1 DD DD RIP1 TRADD TRAF2 Complex I RIP1 RIP1 cIAP1 cIAP2 Dissociation LUBACs TRADD TRAF2 FADD Complex II TAB1 Procaspase-8 (inactive) Caspase-8 (active) Apoptosis TRADD TRAF2 TAB2 TAK1 NF-κB MAP kinase activation cascade cFLIP Survival FIGURE 4-14 Signaling through TNF-R family receptors. Signaling through TNF-R family receptors can lead to pro- or antiapoptotic outcomes depending on the nature of the signal, the receptor, and the cellular context. (a) Apoptosis. Binding of TNF to TNF-R1 induces trimerization of the receptor and conformational alteration in its cytoplasmic domain, resulting in the recruitment of the DD-containing adapter molecule TRADD to the cytoplasmic face of the receptor. TRADD binds to the serine-threonine kinase RIP1 and the TNF receptor associated factor TRAF2. This complex of TRADD, RIP1, and TRAF2, known as complex I, dissociates from the receptor and migrates to the cytoplasm where it binds to the adapter protein FADD. FADD recruits procaspase-8, leading to apoptosis as described in Figure 4-13. The proapoptotic complex generated upon TNF-R1 receptor binding is shown as Complex II. (b). Survival. As for the apoptotic pathway, TNF ligation results in receptor trimerization, TRADD binding, and RIP1 recruitment. In this case, however, TRADD also recruits the ubiquitin ligases cIAP1 and cIAP2, which in turn bind to the proteins of the linear ubiquitin assembly complex (LUBAC) proteins. Polyubiquitination of RIP1 allows it to bind to the NEMO component of the IKK complex as well as to TAK1. TAK1 and RIP1 together activate the IKK complex, leading to IB phosphorylation and destruction, and release of NF-B to enter the nucleus. Among other prosurvival effects, NF-B activates the transcription of the cFLIP protein, which inhibits caspase-8 action, thus tipping the scales in favor of survival. The TAK1 complex also activates MAP kinase signaling, which enhances cell survival. Receptors and Signaling: Cytokines and Chemokines The survival versus death decisions that are made at the level of the TNF-R1 receptor depend upon the outcome of the race between the generation of active caspase-8 on the one hand and the generation of the caspase-8 inhibitor cFLIP on the other. Although we now understand the molecular mechanisms that bring about the consequences of these decisions, we still have much to learn regarding how the cell integrates the signals received through TNF-R1 with other signals delivered to the cell in order to determine which of the two competing pathways will prevail. Since the generation of the membrane-bound complex that is capable of activating NF-␬B is entirely dependent on interactions between various ubiquitinated proteins, it now appears that the life-death decision for a cell may be executed by a small protein previously thought to have only destructive intent. The IL-17 Family Is a Recently Discovered, Proinflammatory Cytokine Cluster The most recently described family of cytokines, the IL-17 family, includes interleukins 17A, 17B, 17C, 17D, and 17F. Signaling through most members of this family culminates in the generation of inflammation. IL-17 receptors are found on neutrophils, keratinocytes, and other nonlymphoid cells. Members of the IL-17 family therefore appear to occupy a TABLE 4-5 | CHAPTER 4 127 location at the interface of innate and adaptive immunity. IL-17 cytokines do not share sequence similarity with other cytokines, but intriguingly the amino acid sequence of IL-17A is 58% identical to an open reading frame (ORF13) found in a T-cell–tropic herpesvirus. The significance of this sequence relationship is so far unknown; did the virus hijack the cytokine sequence for its own needs, or did the pilfering occur in the opposite direction? IL-17 Cytokines IL-17A, the first member of this family to be identified, is released by activated T cells and stimulates the production of factors that signal a proinflammatory state, including IL-6, CXCL8, and granulocyte colony-stimulating factor (G-CSF). As characterization of IL-17A and the T cells that secreted it progressed, it became clear that the T cells secreting this cytokine represent a new lineage, the TH17 cell subset, which is currently the focus of intense investigation (see Chapter 11). Genomic sequencing has since led to the identification of a number of homologs of IL-17A (see Table 4-5). Most of the interleukins in the IL-17 family share the property of operating at the interface of innate and adaptive immunity, serving to coordinate the release of proinflammatory and neutrophil-mobilizing cytokines. However, IL-17E provides an exception to this general rule, instead promoting Expression and known functions of members of the extended IL-17 receptor family Family member Other common names IL-17A Receptors Expression by which cells Main functions IL-17 and CTL-8 IL-17RA and IL-17RC TH17 cells, CD8 T cells,  T cells, NK cells, and NKT cells Autoimmune pathology, neutrophil recruitment, and immunity to extracellular pathogens IL-17B NA IL-17RB Cells of the GI tract, pancreas, and neurons Proinflammatory activities? IL-17C NA IL-17RE Cells of the prostate and fetal kidney Proinflammatory activities? IL-17D NA Unknown Cells of the muscles, brain, heart, lung, pancreas, and adipose tissue Proinflammatory activities? IL-17E IL-25 IL-17RA and IL-17RB Intraepithelial lymphocytes, lung epithelial cells, alveolar macrophages, eosinophils, basophils, NKT cells, TH2 cells, mast cells, and cells of the gastrointestinal tract and uterus Induces TH2 responses and suppresses TH17 responses IL-17F NA IL-17RA and IL-17RC TH17 cells, CD8 T cells,  T cells, NK cells, and NKT cells Neutrophil recruitment and immunity to extracellular pathogens IL-17A/IL-17F heterodimer NA IL-17RA and IL-17RC TH17 cells, CD8 T cells,  T cells, NK cells, and NKT cells Neutrophil recruitment and immunity to extracellular pathogens vIL-17 ORF13 IL-17RA (and IL-17RC?) Herpesvirus saimiri Unknown Adapted from Gaffen, S. L. 2009. Structure and signalling in the IL-17 receptor family. Nature Reviews Immunology 9:556–567. PA R T I 128 IL-17F IL-17A FN1 | IL-17A– IL-17F Introduction IL-17E IL-17B IL-17C IL-17RB IL-17RE IL-17D Unknown ligand Unknown ligand IL-17A FN2 SEFIR TILL CBAD IL-17RB IL-17RC IL-17RA IL-17RA FIGURE 4-15 The IL-17 family of cytokines and their associated receptors. The cytokines that form the IL-17 family share a highly conserved structure, with four conserved cysteines. Only one of the proteins has so far been subject to x-ray analysis, which demonstrates that the structure is that of a “cysteine knot,” a tightly folded protein that exists naturally as a dimer. The five proteins that make up the IL-17 receptor family are IL-17RA, IL-17RB, IL-17RC, IL-17RD, and IL-17RE. These are arranged into homo- and hetero- dimers and trimers to create the complete receptor mol- the differentiation of the anti-inflammatory TH2 subclass, while suppressing further TH17 cell responses, in what amounts to a negative feedback loop. In general, members of the IL-17 family exist as homodimers, but heterodimers of IL-17A and IL-17F have been described. Monomeric units of IL-17 family members range in molecular weight from 17.3 to 22.8 kDa, and crystallographic analysis has revealed that they share a structure that is primarily  sheet in nature, stabilized by intrachain disulfide bonds. The IL-17 Family Receptors The IL-17 receptor family is composed of five protein chains— IL-17RA, IL-17RB, IL-17RC, IL-17RD, and IL-17RE—which are variously arranged into homo- and hetero- dimeric and trimeric units to form the complete receptor molecules (Figure 4-15). Members of the IL-17 receptor family share fibronectin domains with the Hematopoietin and Interferon family cytokine receptors and are single transmembrane proteins. They all contain cytoplasmic SEF/IL-17R (Similar Expression to Fibroblast growth factor interleukin 17 Receptor, or SEFIR) domains, responsible for mediating the protein-protein interactions of the IL-17R signal transduction pathway. The IL17RA chain also contains a TIR-like loop (TILL) domain, analogous to structures found in the Toll and IL-1 receptor molecules, as well as a C/EBP activation domain (CBAD), capable of activating the C/EBP transcription factor. Signaling Through IL-17 Receptors Analogous to signaling through IL-1 receptors, signaling through most IL-17 receptors results in an inflammatory response, and so it should not come as a surprise to learn that signaling through the IL-17 receptor results in activa- Unknown receptor IL-17RD IL-17RD IL-17RA ecules shown. Each receptor protein includes one or more fibronectin (FN) domains, as well as a cytoplasmic SEF/IL-17R (SEFIR) domain that is important in mediating downstream signaling events. The IL-17RA protein also includes a TIR-like loop domain (TILL), similar to that found in Toll-like receptors and IL-1 receptors, as well as a C/EBP activation domain, capable of interacting with the downstream transcription factor C/EBP. [Adapted from S. Gaffen, 2009, Structure and signalling in the IL-17 receptor family, Nature Reviews Immunology, 9:556.] tion of NF-B, a hallmark transcription factor of inflammation. Details of the signaling pathways that emanate from the IL-17R are still being worked out, but Figure 4-16 illustrates the major features of our current knowledge. 1. NF-B activation via IL-17RA and IL-17RC. Binding of IL-17A to the receptor molecules IL-17RA and IL-17RC results in the recruitment of the adapter protein ACT1 to the SEFIR domain. ACT1 binds other proteins, including TRAF3 and TRAF6, which then engage with the TAK1 complex. TAK1 activation results in the phosphorylation and inactivation of the inhibitor of NF-B (IB), allowing NF-B activation and nuclear migration. 2. Activation of MAP kinase pathway and cytokine mRNA stabilization. Adapter proteins bound to the receptor also recruit components of the MAP kinase pathway, resulting in the activation of MAP kinases, including the extracellular signal-regulated kinase Erk1. Though unusual, it appears that the most important role of Erk1 in IL-17 signaling is not in the generation of phosphorylated transcription factors (as is the case for its involvement in TCR- and BCR-mediated cell signaling), but rather in controlling the stability of cytokine mRNA transcripts. Many of the target genes of IL-17 signaling are cytokines and chemokines whose transcription is up-regulated on receipt of an IL-17 signal. The levels of cytokine mRNA are controlled in part by binding of the cytoplasmic protein tristetraprolin to AU-rich elements (AREs) in the 3-untranslated regions of mRNA transcripts. Tristetraprolin then delivers the cytokine mRNAs to the exosome complexes of the cells, where they are degraded. However, phosphorylation of tristetraprolin by MAP kinases inhibits Receptors and Signaling: Cytokines and Chemokines 1 SEFIR ACT1 TRAFs 2 TAB2 MAP kinase cascade 3 TAK1 ERK1 NF-κB activation mRNA stabilization NF-κB C/EBPs Cytokine expression IL-6 expression FIGURE 4-16 Signaling from the IL-17 receptor. Binding of IL-17 to its receptor initiates three signaling pathways. (1) Binding of IL-17 to its receptor results in the recruitment of the adapter protein ACT1 to the cytoplasmic region of the receptor. ACT1 then serves as a docking point for TRAF proteins 3 and 6, which in turn recruit members of the TAK1 complex, consisting of the TAK1 kinase and TAK1 binding proteins. TAK1 activation results in the phosphorylation and activation of the IKK complex and resultant NF-B activation, as described previously. (2) Adapter proteins bound to the SEFIR (Similar Expression to Fibroblast growth factor interleukin 17 Receptor) domain also recruit components of the MAP kinase pathway. The MAP kinase Erk1 phosphorylates the cytoplasmic protein tristetraprolin, and inhibits its ability to bind to AU-rich elements on mRNA encoding cytokines. Since tristetraprolin binding results in mRNA degradation, activation of this arm of the pathway results in enhancing the stability of cytokine mRNA. (3) IL-17 binding to its receptor also results in the activation of transcription factors of the C/EBP family, which promote the expression of the inflammatory cytokine IL-6. its ability to recruit the degradative machinery and hence results in increased stability of the cytokine and chemokine mRNAs. 3. CHAPTER 4 129 Chemokines Direct the Migration of Leukocytes Through the Body IL-17R TAB1 | Activation of the transcription factors C/EBP and C/EBP In addition to up-regulation of NF-B and members of the MAP kinase pathway, signaling through IL-17RA and IL-17RC proteins has also been shown to activate the transcription factors C/EBP and C/EBP , which promote expression of IL-6, one of the quintessential inflammatory cytokines. Chemokines are a structurally related family of small cytokines that bind to cell-surface receptors and induce the movement of leukocytes up a concentration gradient and toward the chemokine source. This soluble factor-directed cell movement is known as chemotaxis, and molecules that can elicit such movement are referred to as chemoattractants (Box 4-3). Some chemokines display innate affinity for the carbohydrates named glycosaminoglycans, located on the surfaces of endothelial cells, a property that enables them to bind to the inner surfaces of blood vessels and set up a cellbound chemoattractant gradient along blood vessel walls, directing leukocyte movement. Chemokine Structure Chemokines are relatively low in molecular weight (7.5– 12.5kDa) and structurally homologous. The tertiary structure of chemokines is constrained by a set of highly conserved disulfide bonds; the positions of the cysteine residues determine the classification of the chemokines into six different structural categories (Figure 4-17). Within any one category, chemokines may share 30% to 99% sequence identity. The grouping of chemokines into the subclasses shown in Figure 4-17 has functional, as well as structural, significance. For example, the seven human CXC chemokines within the ELR subclass share the same receptor (CXCR2), attract neutrophils, are angiogenic, and have greater than 40% sequence identity. (A substance is angiogenic if it promotes the formation of new blood vessels; it is angiostatic if it prevents the formation of new blood vessels.) The non-ELR, CXCL chemokines CXCL9, CXCL10, and CXCL11, are also more than 40% identical to one another; however, this group is angiostatic, not angiogenic, and utilizes the CXCR4 receptor. Members of the two, structurally distinct CC groups are chemoattractants that attract monocytes and macrophages (although not neutrophils) to the site of infection. See Appendix III for a more comprehensive tabulation of chemokines and their immunologic roles. Chemokine Receptors In the 1950s, investigations of the mechanisms by which glucagon and adrenaline signaling led to an increase in the rate of glycogen metabolism revealed the existence of a class of receptors that threads through the membrane seven times and transduces the ligand signal via interactions with a polymeric GTP/GDP-binding “G protein.” This class of G-Protein– Coupled Receptors (GPCRs) is used in the recognition of many types of signals, including those mediated by chemokines. Certain essential features of this pathway are conserved in all GPCR-type responses. (These larger, polymeric, seven membrane pass receptor-associated G proteins are different from the small, monomeric G proteins such as ras, which participate farther downstream in intracellular signaling pathways. Although both types of G proteins are 130 PA R T I | Introduction ADVANCES How Does Chemokine Binding to a Cell-Surface Receptor Result in Cellular Movement Along the Chemokine Gradient? Chemotaxis is the mechanism by which the speed and direction of cell movement are controlled by a concentration gradient of signaling molecules. Receptors on the surface of responsive cells bind to chemotactic factors and direct the cell to move toward the source of factor secretion. Chemotaxis was recognized as a biological phenomenon as early as the 1880s, and is found in organisms as simple as bacteria, which use chemotaxis to locate sources of nutrition. However, only as scientists have developed the ability to analyze signaling pathways and to observe the movement of individual cells under in vivo conditions have we been able to approach a mechanistic understanding of the complex series of intracellular events that culminate in chemically directed cell movement in the immune system. Although we still do not understand all the details of this process, some general principles have begun to emerge. We first address the means by which cells sense chemoattractant signals. We next develop an appreciation for how cells move, and finally we can describe a little of what we know regarding how they become polarized in order to direct their movement as specified by the chemotactic signal. SIGNAL SENSING Leukocytes recognize chemoattractant signals, or chemokines, using the G-protein– coupled receptors described in this chapter. Such receptors are located at the plasma membrane, and the cytoplasmic face of these receptors is associated with a polymeric G protein, so called because of its affinity for guanosine phosphates. On binding to its chemokine ligand, the receptor alters its conformation, passing on that conformational change to its associated G protein. The bound G protein then loses affinity for guanosine diphosphate (GDP) and instead binds guanosine triphosphate (GTP), thus achieving its active conformation. Next, the G protein dissociates into two subunits, termed G-GTP and G. The signaling pathways emanating from each of these subunits are described in the text. It is difficult to overstate the effectiveness of these receptor signals in determining the direction and speed of cellular movement. For example, neutrophils can recognize and move in response to extremely shallow chemoattractant gradients, in which the concentration of the chemoattractant at the front of the gradient is as little as 2% higher than that at the back. and almost like wires (filopodia), or protrude from the surface of the cell like a sheet (lamellipodia). The nucleus of the migrating cell lies behind these protrusions in the cell body, and at the posterior part of the cell lies a near-cylindrical tail, the uropod, which may be as long as 10 m. Formation of protrusions Protrusions Direction of movement Formation of adhesions Uropod Adhesions Contraction HOW CELLS MOVE Figure 1 illustrates a model of amoeboid cellular movement, the mechanism used by leukocytes. Ameboid migration is a rapid type of cell movement; leukocytes and stem cells, which use this mode of migration, may move as quickly as 30 m/ minute. Cells routinely move over other cells and tissues (the substrata) by first forming cellular protrusions at their leading edge—the edge closest to the direction of movement. Depending on the cell type and the nature of the chemoattractant signal, these protrusions may be thin, FIGURE 1 Schematic description of the classical migratory cycle. [Adapted from P. Mrass et al., 2010, Cell-autonomous and environmental contributions to the interstitial migration of T cells, Seminars in Immunopathology 32:257.] Receptors and Signaling: Cytokines and Chemokines | CHAPTER 4 131 BOX 4-3 The protrusions at the leading edge of a moving leukocyte form noncovalent attachments to the substratum using adhesive interactions between proteins on the cell protrusion and other proteins on the substratum. The substratum may, for example, be a capillary endothelial cell, when a white blood cell rolls along the inside of a blood capillary. Alternatively, during leukocyte movement within a lymph node, the adhesive interaction may occur between B or T lymphocyte surface integrins and proteins located on a cellular extension from a follicular reticular cell. Once the migrating cell is temporarily attached to the substratum, contractions of the cell body mediated by actin and myosin, and resulting cytoskeletal rearrangements, bring the rest of the cell forward toward the leading edge and then the cycle of reaching forward with cellular protrusions and contraction of the rest of the cell in the direction of the leading edge can recur. GENERATION OF POLARITY SIGNALS: HOW CELLS DECIDE ON THE DIRECTION OF MOVEMENT FIGURE 2 Polarization of neutrophils after receipt of a chemoattractant signal. A neutrophil exposed to a chemoattractant organizes filamentous actin at the protrusions on the leading edge of the cell (shown here in red) and an actino-myosin contractile complex at the uropod and along the sides of the cell (shown here stained green). [Fei Wang, The signaling mechanisms underlying cell polarity and chemotaxis. ColdSpring Harbor Perspectives in Biology, October 2009, 1(4):a002980. doi: 10.1101/cshperspect.a002980. ? 2009 Cold Spring Harbor Laboratory Press, all rights reserved.] Upon receipt of a chemotactic signal, leukocytes polarize very rapidly, and leukocyte locomotion is clearly visible within 30 seconds after the receipt of the chemokine by the cell. Scientists initially hypothesized that movement of a leukocyte toward a particular chemoattractant may result from an asymmetric distribution of chemoattractant receptors on the cell surface, with a higher concentration of receptors localized to the leading edge of the cell. However, microscopic observations suggest that this initial hypothesis was incorrect. Instead, investigators found an asymmetrical distribution of signaling molecules and cytoskeletal components (Figure 2). Thus, although there is a uniform distribution of receptors on the cell surface, the distribution of occupied receptors is asymmetrical, which means that the signal received from chemokines on the side of the cell facing the chemokine source is stronger than that received elsewhere on the cell surface. The ultimate result of this asymmetrical signal from the occupied receptors is an internal redistribution of signaling pathway components and cytoskeletal elements; this in turn causes the cell to move toward the source of the signal. Many of the signaling pathways that mediate this directional movement are already familiar to you from Chapter 3 and from this chapter. The  dimer of the activated G protein recruits a particular subclass of PI3 kinase enzymes to the inner leaflet of the leading edge of the membrane, where it phosphorylates PIP2 and other phosphatidyl inositol phosphates. These phosphorylated lipids then serve as docking sites for PHdomain–containing proteins, including Akt, which in turn phosphorylates downstream effectors leading to actin polymerization. Depending on the cell type and the nature of the stimulation, the small GTPases Rac, Cdc42, and Rho are also implicated in the chemokinemediated cytoskeletal modifi cations that are necessary for cell movement, and these are activated via tyrosine kinases in lymphocytes and by pathways that are as yet not fully characterized in other leukocyte cell types. A cell can only move in response to a gradient if the trailing edge of the cell moves at the same rate as the leading edge, and the uropod of moving leukocytes is susceptible to its own set of G-protein–mediated signals. Following redistribution of the internal signaling components, the chemokine receptors at the trailing edge of the cell are coupled to a set of G-protein trimers different from those at the leading edge. Interaction of the uropod-localized G-Protein–Coupled Receptors (GPCRs) leads to the formation and contraction of actin-myosin complexes and subsequent retraction of the uropod. 132 PA R T I | Introduction Class Structural signature Names Number (n) in class CX3C .............CXXXC.............C.........C......... CX3CL1 1 Non-ELR CXC .............CX__C.............C.........C......... CXCL# 9 ELR CXC ...ELR....CX__C.............C.........C......... CXCL# 7 4C CC .............C___C.............C.........C......... CCL# 19 6C CC .............C___C......C.....C.........C....C... CCL# 5 C .....................C........................C XCL# 2 FIGURE 4-17 Disulfide bridges in chemokine structures. A schematic of the locations of cysteine residues in chemokines that shows how the locations of cysteines determine chemokine class. Chemokines are proteins of small molecular weight which share two, four, or six conserved cysteine residues at particular points in their sequence that form intrachain disulfide bonds. The number of cyste- activated by GTP binding, their structures and functions are quite different). The GPCRs are classified according to the type of chemokine they bind. For example, the CC receptors (CCRs) recognize CC chemokines, the CXCRs recognize CXCL chemokines, and so on. Chemokine receptors bind to their respective ligands quite tightly (Kd ≅ 109 M). Interestingly, the intrinsic specificity of the receptors is balanced by the capacity of many receptors to bind more than one chemokine from a particular family and of several chemokines to bind to more than one receptor. For example, the receptor CXCR2 recognizes seven different chemokines, and CCL5 can bind to both CCR3 and CCR5. Signaling Through Chemokine Receptors The cytoplasmic faces of seven membrane pass GPCRs associate with intracellular, trimeric GTP-binding proteins consisting of G, G, and G subunits (Figure 4-18a). When the receptor site is unoccupied, the G subunit of the trimeric G protein binds to GDP. Chemokine ligation to the receptor results in a conformational change that is transmitted to the G protein and in turn induces an exchange of GDP for GTP at the G binding site that is analogous to what occurs upon GTP binding to the small GTP-binding proteins such as Ras. This results in the dissociation of the G protein into a G-GTP monomer and a G dimer. In the case of the chemokine receptor associated G proteins, chemokine signaling is mediated by both the dissociated G dimer (which has no nucleotide binding site) and the G-GTP subunit. Just as for the small G-protein–coupled pathways, the duration of signaling through the chemokine receptor is limited by the intrinsic GTPase activity of the G subunit, which in turn can be increased by GTPase Activating ines as well as the positions of the disulfide bonds determine the subclass of these cytokines as shown. The overscores indicate the cysteines between which disulfide bonds are made. The naming of chemokines in part reflects the cysteine-determined class (see Appendix III). [Adapted from W. E. Paul, 2008, Fundamental Immunology, 6th ed. Lippincott Williams & Wilkins, Philadelphia, Figure 26.1] Proteins (GAPs), also known as Regulators of G-protein Signaling (RGSs). Because the pathway is active only when the protein binds to GTP, GAPs down-regulate the activity mediated by the receptor (see Figure 3-15). Once the GTP in the G binding site is hydrolyzed to GDP, G re-associates with the G dimers, effectively terminating signaling. There are multiple subtypes of G and G subunits that vary in representation between different cell types. Signaling through different subunits can give rise to different consequences, depending on the downstream pathways that are elicited. Just as for the small G protein, there are several different large polymeric G proteins, which vary in their cellular distribution and receptor partners. Once released from its G partner, the G subunit of the trimeric G protein activates a variety of downstream effector molecules, including those of the Ras/MAP kinase pathway (path 1 in Figure 4-18). Full activation of MAP kinase is further facilitated by tyrosine phosphorylation mediated by G-GTP-activated tyrosine kinases (not shown). Activation of the Ras pathway culminates in the initiation of transcription as well as in up-regulation of integrin adhesion molecules on the cell membrane. G signaling also cooperates with signaling mediated by G-GTP to activate one isoform of phospholipase C, PLC, resulting in an increase in the activity of the transcription factor NF-B (path 2). The GGTP complex also activates a signaling pathway that is initiated by the small G protein, Rho (path 3). This pathway leads to actin polymerization and the promotion of cell migration, so it is this third pathway that is responsible for the most commonly described aspect of chemokine signaling: cell movement (see also Box 4-3). Rho signaling is also instrumental in bringing about changes in the transcriptional program of the cell. Finally, a JAK associated with Receptors and Signaling: Cytokines and Chemokines Chemokine Chemokine receptor C JAK 4 β γ α 2 PLCβ G protein 3 1 PKC Ras Rho Cell movement Akt MAP kinase cascade Survival | CHAPTER 4 133 the GPCR initiates signaling through PKC-mediated pathways that culminate in the activation of Akt and increased cell survival (path 4) as well as in further transcriptional alterations. The quality of the response elicited by particular chemokines in different types of cells is dependent on the nature of the chemokine ligand, as well as on the signaling microenvironment, which in turn is determined by the range of G protein subtypes and regulatory molecules present in that cell. But from the complexity of signaling options available to the receptor, it is not difficult to see how binding of a chemokine molecule to its receptor can simultaneously bring about alterations in the location, the adhesion molecule binding capacity, and the transcriptional program of the chemokine-activated cell. Cytokine Antagonists AP-1 NF-κB Gene expression FIGURE 4-18 G-protein–coupled receptors interact with G proteins that transduce chemokine signals into the interior of the cell. Different chemokines can induce different signaling pathways. This figure therefore represents a composite of some of the most common pathways elicited by chemokine binding, which lead collectively to alterations in the transcriptional program, the enabling of cell movement, and changes in the adhesive properties of the signaled cell. (1) The G subunit binds to the adapter molecule Grb2, activating it and initiating the Ras signaling pathway that leads eventually to activation of MAP kinase and an alteration in the cell’s transcriptional program, as shown in Figure 3-16. Ras pathway activation also leads eventually to activation of integrin adhesion molecules on the cell surface. G GTP simultaneously binds and activates a protein tyrosine kinase that phosphorylates and further activates MAP kinase (not shown). (2) Both GGTP and G cooperate to activate PLC, which activates the NF-B pathway. (3) GGTP activates the small cytoplasmic G protein, Rho, initiating actin polymerization and cell movement. Other pathways emanate from Rho that lead to the activation of the transcription factor Serum Response Factor (SRF). (4) A JAK is stimulated by chemokine binding to the receptor, and it turns on the activity of PKC, leading eventually to the activation of the enzyme Akt. Akt affects cell survival by phosphorylating the proapoptotic genes Bax and Bad (not shown) and marking them for destruction and enhancing cell survival. It also phosphorylates and further activates the transcription factor NF-B. JAK-mediated PKC activation can also lead to phosphorylation of the transcription factor Jun, and its dimerization with Fos to create the complete transcription factor AP-1. A number of proteins that inhibit the biological activity of cytokines have been reported. These proteins act in one of two ways: either they bind directly to a cytokine receptor but fail to activate the cell, thus blocking the active cytokine from binding, or they bind directly to the cytokine itself, inhibiting its ability to bind to the cognate receptor. In this section, we describe some naturally occurring cytokine antagonists that modulate and refine the power of particular cytokine responses, as well as the ways in which various pathogens have hijacked cytokine responses to their own ends. The IL-1 Receptor Antagonist Blocks the IL-1 Cytokine Receptor The best-characterized cytokine inhibitor is the IL-1 receptor antagonist (IL-1Ra), which binds to the IL-1 receptor but does not elicit activation of the signaling pathway (see above). As previously described, ligation of IL-1Ra to the IL-1 receptor blocks the binding of both IL-1 and IL-1, thus accounting for its inhibitory properties. IL-1Ra is synthesized by the same cells that secrete IL-1 and IL-1, and its synthesis in the liver is up-regulated under inflammatory conditions, along with that of IL-1. Several animal and human models exist in which the levels of IL-1Ra are naturally reduced, and humans carrying an allele that decreases the expression of IL-1Ra suffer from arthritis and a variety of other autoimmune diseases. This observation suggests that the normal function of IL-1Ra is to provide for the host a means by which to modulate the numbers of receptors that are capable of mounting a physiological response to IL-1. Given the fiercely proinflammatory effects of IL-1, it makes sense that responses to this powerful cytokine should be carefully controlled. Indeed, we note quite often in biological systems that a process—if it has the potential to lead to deleterious consequences to the organism—is subject to several different means of regulation. 134 PA R T I | Introduction Recombinant IL-Ra has been used clinically, under the name of anakrina, for the treatment of rheumatoid arthritis. Investigations into how cells control the balance of IL-1 and IL-1Ra secretion are still ongoing, but preliminary studies suggest that the activation of different isoforms of PI3 kinase may play an important role in determining the relative amounts of IL-1 and IL-1Ra that are secreted by a stimulated monocyte. Cytokine Antagonists Can Be Derived from Cleavage of the Cytokine Receptor Some naturally occurring soluble antagonists arise from enzymatic cleavage of the extracellular domains of cytokine receptors. These soluble receptor components can compete with the membrane-bound receptor for cytokine binding and thus down-modulate the potential cytokine response. The best characterized of the soluble cytokine receptors consists of a segment containing the amino-terminal 192 amino acids of the IL-2R (CD25) subunit, which is released by proteolytic cleavage, forming a 45-kDa soluble IL-2 receptor (sIL-2R or sCD25). The shed receptor retains its ability to bind IL-2 and can therefore prevent the cytokine’s productive interaction with the membrane-bound IL-2 receptor. The origin of sIL-2R is still a matter for debate. Recently, regulatory T cells, which express high levels of CD25 on their membrane surfaces, have been shown to release sIL-2R upon activation. Since these T cells serve the function of down-regulating ongoing immune responses, it has been suggested that the soluble IL-2 receptors may serve the physiological function of soaking up excess IL-2 and thus reducing the amount of the cytokine that is available to irrelevant or even to competing effector T cells. The soluble IL-2 receptor is also found in the serum and bodily fluids of patients suffering from a number of hematologic malignancies (blood cell cancers), and high levels of sIL-2R in the blood correlate with a poor disease prognosis. However, the issue of whether the sIL-2R is released from the tumor cells per se, or whether it is released from regulatory T cells that may be acting to dampen the host anti-tumor response, has yet to be resolved. Some Viruses Have Developed Strategies to Exploit Cytokine Activity The cytokine antagonists described above derive from an organism’s own immune system. However, as is so often the case in immunology, some pathogens have evolved ways in which to circumvent cytokine responses, by mimicking molecules and pathways used by the host. The evolution of anticytokine strategies by microbial pathogens provides biological evidence of the importance of cytokines in organizing and promoting effective antimicrobial immune responses. The following are among the various anti-cytokine strategies used by viruses: • The generation of viral products that interfere with cytokine secretion • The generation of cytokine homologs that compete with natural cytokines or inhibit anti-viral responses • The production of soluble cytokine-binding proteins • The expression of homologs of cytokine receptors • The generation of viral products that interfere with intracellular signaling • The induction of cytokine inhibitors in the host cell Epstein-Barr virus (EBV), for example, produces an IL-10–like molecule (viral IL-10 or vIL-10) that binds to the IL-10 receptor. Just like host-derived IL-10, this viral homologue suppresses TH1-type cell-mediated responses that would otherwise be effective in fighting a viral infection. Other cytokine mimics produced by viruses allow them to manipulate the immune response in alternative ways that aid the survival of the pathogen. For example, EBV produces an inducer of IL-1Ra, the host antagonist of IL-1. Poxviruses have also been shown to encode a soluble TNF-binding protein and a soluble IL-1-binding protein that block the ability of the bound cytokines to elicit a response. Since both TNF- and IL-1 are critical to the early phases of an inflammatory, antiviral response, these soluble cytokine-binding proteins may allow the viruses an increased time window in which to replicate. Yet other viruses produce molecules that inhibit the production of cytokines. One such example is the cytokine response modifier (Crm) protein of the cowpox virus, which inhibits the production of caspase-1 and hence prevents the processing of IL-1 precursor proteins. Finally, some viruses produce soluble chemokines and chemokine-binding proteins that interfere with normal immune cell trafficking, and allow the producing viruses and virally infected cells to evade an immune response. Table 4-6 lists a number of viral products that inhibit cytokines, chemokines, and their activities. Cytokine-Related Diseases Defects in the complex regulatory networks governing the expression of cytokines and cytokine receptors have been implicated in a number of diseases. Genetic defects in cytokines, their receptors, or the molecules involved in cytokinedirected signal transduction lead to immunodeficiencies such as those described in Chapter 18. Other defects in the cytokine network can cause an inability to defend against specific families of pathogens. For example, people with a defective receptor for IFN- are susceptible to mycobacterial Receptors and Signaling: Cytokines and Chemokines TABLE 4-6 | CHAPTER 4 135 Viruses use many different strategies to evade cytokine-mediated immune mechanisms Virus Virally encoded proteins Epstein-Barr Virus (EBV), Cytomegalovirus IL-10 homolog Vaccinia virus, Variola virus Soluble IL-1 receptors Myxoma virus Soluble IFN- receptor Variola virus Soluble TNF receptors Adenovirus RID complex proteins induce internalization of Fas receptor Measles virus Viral hemagglutinin binds to complement receptor, CD46, signaling disruption of IL-12 production and therefore inhibition of TH1 pathway differentiation Herpes simplex virus Reverses translation block induced by Type 1 interferons Adenovirus Blocks interferon-induced JAK/STAT signaling infections that rarely occur in the general population. In addition to the diseases rooted in genetic defects in cytokine activity, a number of other pathologic states result from overexpression or underexpression of cytokines or cytokine receptors. Several examples of these diseases are given below, followed by an account of therapies aimed at preventing the potential harm caused by cytokine activity. Septic Shock Is Relatively Common and Potentially Lethal Despite the widespread use of antibiotics, bacterial infections remain a major cause of septic shock, which may develop a few hours after infection by certain bacteria, including Staphyloccocus aureus, E. coli, Klebsiella pneumoniae, Pseudomonas aeruginosa, Enterobacter aerogenes, and Neisseria meningitidis. Bacterial septic shock is one of the conditions that falls under the general heading of sepsis. Sepsis, in turn, may be caused not only by bacterial infection but also by trauma, injury, ischemia (decrease in blood supply to an organ or a tissue), and certain cancers. Sepsis is the most common cause of death in U.S. hospital intensive-care units and the 13th leading cause of death in the United States. A common feature of sepsis, whatever the underlying cause, is an overwhelming production of proinflammatory and fever-inducing cytokines such as TNF- and IL-1. The cytokine imbalance induces abnormal body temperature, alterations in the respiratory rate, and high white blood cell counts, followed by capillary leakage, tissue injury, widespread blood clotting, and lethal organ failure. Bacterial septic shock often develops because bacterial cell wall endotoxins bind to innate immune system pathogen receptors, such as Toll-like receptors (see Chapter 5), on dendritic cells and macrophages, causing them to produce IL-1 and TNF- at levels that lead to pathological capillary perme- ability and loss of blood pressure. A condition resembling bacterial septic shock can be produced in the absence of any bacterial infection simply by injecting mice with recombinant TNF-. Several studies offer hope that neutralizing TNF- or IL-1 activity with monoclonal antibodies or antagonists will prevent fatal shock from developing. In one such investigation, monoclonal antibodies to TNF- protected animals from endotoxin-induced shock. In another, injection of a recombinant IL-1 receptor antagonist (IL-1Ra), which prevents binding of IL-1 to the IL-1 receptor as described above, resulted in a threefold reduction in mortality. However, neutralization of TNF- does not reverse the progression of septic shock in all cases, and antibodies against TNF- give little benefit to patients with advanced disease. Recent studies in which the cytokine profiles of patients with septic shock were followed over time shed some light on this apparent paradox. The increases in TNF- and IL-1 occur rapidly in early sepsis, so neutralizing these cytokines is most beneficial early in the process. Indeed, in animal experiments, early intervention can prevent sepsis altogether. However, approximately 24 hours following the onset of sepsis, the levels of TNF- and IL-1 fall dramatically, and other factors become more important. Cytokines critical in the later stages of sepsis include IL-6, MIF, and CCL-8. Sepsis remains an area of intense investigation, and clarification of the process involved in bacterial septic shock and other forms of sepsis can be expected to lead to advances in therapies for this major killer in the near future. Bacterial Toxic Shock Is Caused by Superantigen Induction of T-Cell Cytokine Secretion A variety of microorganisms produce toxins that act as superantigens. As described in Chapters 9 and 11, superantigens 136 PA R T I | Introduction BOX 4-4 CLINICAL FOCUS Cytokines and Obesity Even as those in the Third World suffer from malnutrition, among the leading current causes of disease and death in developed countries are obesity and its corollary, Type 2 diabetes. Type 1, or juvenile onset, diabetes has long been known to have an autoimmune etiology (cause). Cells of the adaptive immune system kill the  cells of the Islets of Langerhans in the pancreas, leading to the complete absence of insulin in the diabetic patient, who must rely on exogenously delivered hormone to survive. In Type 2 diabetes, the body fails to respond to insulin in an appropriate manner, and the cells fail to take in glucose from the blood and into the tissues. Just as in Type 1 diabetes, the presence of high levels of glucose in the blood and tissue fluids facilitates nonenzymatic glucose conversions to reactive carbohydrate derivatives such as glyoxals. These derivatives cross-link proteins and carbohydrates in the membranes and walls of blood vessels and neurons and within the extracellular matrix, leading to the familiar array of diabetic symptoms: poor peripheral circulation, plaque buildup in the arteries, and heart disease. However, we are now beginning to understand that Type 2, just like Type 1, diabetes is a disease closely related to the workings of the immune system. In 1993, a seminal publication by Hotamisligil and colleagues made the link between inflammation and metabolic conditions such as Type 2 diabetes and obesity. These authors demonstrated that adipocytes (fat cells) constitutively express the proinflammatory cytokine TNF-, and that TNF- expression in adipocytes of obese animals is markedly increased. Later research demonstrated that this finding could be extended to humans. The adipose tissue of human subjects was found to constitutively express TNF-, and blood levels of TNF- fall after weight loss. But is this increase in TNF- expression affecting insulin sensitivity? On binding of insulin to its receptor, tyrosine kinase activity in the cytoplasmic region of the receptor is activated. This Receptor Tyrosine Kinase (RTK) then phosphorylates both itself (autophosphorylation) and some nearby proteins. The signaling cascade is initiated from the insulin receptor by the binding of adapter proteins such as Grb2 and IRS-1 via their SH2 domains to the phosphorylated tyrosine sites on the receptor molecule. We now know that TNF- signaling in adipocytes inhibits the autophosphorylation of tyrosine residues of the insulin receptor and instead induces phosphorylation of serine residues of both the insulin receptor and the IRS-1 adapter. The serine phosphorylation inhibits any subsequent phosphorylation at tyrosine residues, and thus the passage bind to MHC Class II molecules at a location in the MHC molecule that is outside the groove normally occupied by antigenic peptides (see Figure 11-6). They then bind to a part of the V chain of the T-cell receptor that is outside the normal antigen-binding site, and this binding is sufficient to trigger T-cell activation. This means that a given superantigen can simultaneously activate all T cells bearing a particular V domain. Because of their unique binding ability, superantigens can activate large numbers of T cells irrespective of the antigenic specificity of their canonical antigenbinding site. of signals from insulin to the interior of the fat cell is prevented. Recently, the interleukin IL-6 has also been shown to inhibit insulin signal transduction in hepatocytes (liver cells) through a similar mechanism. The decrease in effectiveness of insulin signaling then becomes a self-reinforcing problem, as insulin signaling itself is antiinflammatory, and so any decrease in the insulin signal can give rise to inflammatory side effects. However, TNF- and IL-6 are not the only cytokines implicated in the etiology of Type 2 diabetes. With the discovery and characterization of the proinflammatory cytokine family represented by IL-17, interest has arisen in the relationship between members of this family, obesity, and the control of fat cell metabolism. Since IL-6 is implicated in the differentiation of T lymphocytes to secrete IL-17, obesity and its associated inflammation tend to predispose an individual to secrete IL-17. However, again we find ourselves in a positive feedback loop, as IL-17 acts on monocytes to induce the further secretion of IL-6, thus ensuring the maintenance of an inflammatory state. It is therefore clear that cytokine signaling plays a profound role in a disease that is emblematic of our time, and which is predicted to afflict close to 40% of the U.S. population by the middle of the next decade. Although less than 0.01% of T cells respond to a given conventional antigen (see Chapter 11), 5% or more of the T-cell population can respond to a given superantigen. Bacterial superantigens have been implicated as the causative agent of several diseases, such as bacterial toxic shock and food poisoning. Included among these bacterial superantigens are several enterotoxins, exfoliating toxins, and toxic shock syndrome toxin (TSST1) from Staphylococcus aureus; pyrogenic exotoxins from Streptococcus pyrogenes; and Mycoplasma arthritidis supernatant (MAS). The large number of T cells activated by these superantigens Receptors and Signaling: Cytokines and Chemokines results in excessive production of cytokines. The TSST1, for example, has been shown to induce extremely high levels of TNF- and IL-1. As in bacterial septic shock, these elevated concentrations of cytokines can induce systemic reactions that include fever, widespread blood clotting, and shock. In addition to those diseases described above, in which cytokines or their receptors are directly implicated, recent information has indicated the importance of cytokine involvement in the most important public health crisis currently afflicting the developed world: the increasing incidence of Type 2 diabetes. The roles of TNF- and IL-6 in the induction and maintenance of this disease are described in Box 4-4. Cytokine Activity Is Implicated in Lymphoid and Myeloid Cancers Abnormalities in the production of cytokines or their receptors have been associated with some types of cancer. For example, abnormally high levels of IL-6 are secreted by cardiac myxoma (a benign heart tumor), myeloma and plasmacytoma cells, as well as cervical and bladder cancer cells. In myeloma and plasmacytoma cells, IL-6 appears to operate in an autocrine manner to stimulate cell proliferation. When monoclonal antibodies to IL-6 are added to in vitro cultures of myeloma cells, their growth is inhibited. In contrast, transgenic mice that express high levels of IL-6 have been found to exhibit a massive, fatal, plasmacell proliferation, called plasmacytosis. In addition, as described above, high serum concentrations of the sIL-2R are found in patients suffering from various blood cell cancers, which may impede a vigorous anti-tumor response. Cytokine Storms May Have Caused Many Deaths in the 1918 Spanish Influenza Occasionally, a particularly virulent infection may induce the secretion of extremely high levels of cytokines, that then feed back on the immune cells to elicit yet more cytokines. Normally, these positive feedback loops represent effective modes of immune amplification; they are themselves usually kept in check by self-regulating immune mechanisms, such as the activation of regulatory T cells (see Chapter 11). However, some viruses cause a localized, exaggerated response, resulting in the secretion of extraordinarily high levels of cytokines. If this occurs in the lungs, for example, the localized swelling, inflammation, and increase in capillary permeability can lead to the accumulation of fluids and leukocytes that block the airways, thereby causing exacerbation of symptoms, or even death, before the cytokine levels can be controlled. It is unclear why some viruses induce these cytokine storms and others do not. | CHAPTER 4 137 Historical documents detailing the symptoms of the 1918 Spanish influenza suggest that the massive fatalities associated with that pandemic most likely resulted from cytokine storms, and there is some evidence that the severe acute respiratory syndrome (SARS) epidemic of 1993 may have caused a similar, unregulated, immune cell cytokine secretion. Transplant surgeons also observe this phenomenon on occasions when leukocytes associated with a graft—often a bone marrow transplant—mount an immune response against the host. Effective treatments for patients undergoing cytokine storms are still being developed. Currently, patients are offered steroidal and nonsteroidal anti-inflammatory medications, but other drugs that are more specifically directed at the reduction of cytokine secretion and/or activity are being tested. Cytokine-Based Therapies The availability of purified cloned cytokines, monoclonal antibodies directed against cytokines, and soluble cytokine receptors offers the prospect of specific clinical therapies to modulate the immune response. Cytokines such as interferons (see Clinical Focus Box 4-2), colony-stimulating factors such as G-CSF, and IL-2 have all been used clinically. In addition, several reagents that specifically block the proinflammatory effects of TNF- have proven to be therapeutically useful in certain diseases. Specifically, soluble TNF- receptor (Enbrel) and monoclonal antibodies against TNF- (Remicade and Humira) have been used to treat rheumatoid arthritis and ankylosing spondylitis in more than a million patients. These anti–TNF- drugs reduce proinflammatory cytokine cascades; help to alleviate pain, stiffness, and joint swelling; and promote healing and tissue repair. In addition, as described above, the recombinant form of IL-1Ra— anakinra (Kineret)—has been shown to be relatively effective in the treatment of rheumatoid arthritis. Monoclonal antibodies directed against the  chain of the IL-2R— basiliximab (Simulect) and daclizumab (Zenapax)—are also in clinical use for the prevention of transplantation rejection reactions. As powerful as these reagents may be, interfering with the normal course of the immune response is not without its own intrinsic hazards. Reduced cytokine activity brings with it an increased risk of infection and malignancy, and the frequency of lymphoma is higher in patients who are long-term users of the first generation of TNF- blocking drugs. In addition, the technical problems encountered in adapting cytokines for safe, routine medical use are far from trivial. As described above, during an immune response, interacting cells may produce extremely high local concentrations of cytokines in the vicinity of target cells, but achieving such high concentrations over a clinically significant time period, when cytokines must be 138 PA R T I | Introduction administered systemically, is difficult. Furthermore, many cytokines have a very short half-life—recombinant human IL-2 has a half-life of only 7 to 10 minutes when administered intravenously—so frequent administration may be required. Finally, cytokines are extremely potent biological response modifiers, and they can cause unpredictable and undesirable side effects. The side effects from administration of recombinant IL-2, for example, range from mild (e.g., fever, chills, diarrhea, and weight gain) to serious (e.g., anemia, thrombocytopenia, shock, respiratory distress, and coma). The use of cytokines and anti-cytokine therapies in clinical medicine holds great promise, and efforts to develop safe and effective cytokine-related strategies continue, particularly in those areas of medicine that have so far been resistant to more conventional approaches, such as inflammation, cancer, organ transplantation, and chronic allergic disease. S U M M A R Y ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ Cytokines are proteins that mediate the effector functions of the immune system. Most cytokines are soluble proteins, but some—for example, members of the TNF family, may be expressed in a membrane-bound form. Some cytokines are secreted following stimulation of the innate immune system (e.g., IL-1, TNF-, CXCL8), whereas others are secreted by the T and B lymphocytes of the adaptive immune system (IL-2, IL-4, IL-17). Cytokines bind to receptors on the plasma membrane and elicit their effects through the activation of an intracellular signaling cascade. Cytokines can effect alterations in the differentiative, proliferative, and survival capacities of their target cells. Cytokines exhibit the properties of redundancy, pleiotropy, synergy, antagonism, and cascade induction. The levels of expression of cytokine receptors on the cell surface may change according to the activation status of a cell. There are six families of cytokines with associated receptors, distinguished on the basis of the structures of the cytokines and the receptor molecules, and on the nature of their signaling pathways. IL-1 family members interact with dimeric receptors to induce responses that are primarily proinflammatory. The physiological responses to some IL-1 family members are modulated by the presence of soluble forms of the receptors and soluble cytokine-binding proteins. The Hematopoietin (Class I cytokine) family is the largest family of cytokines, and members mediate diverse effects, including proliferation, differentiation, and antibody secretion. The Hematopoietin family members share a common, four-helix bundle structure. ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ Receptors for cytokines from the Hematopoietin family are classified into three subgroups—the , , or gp130 receptors—each of which shares a common signaling chain. The Interferon (Class II cytokine) family includes the Type I interferons (interferon  and interferon ), which were the first cytokines to be discovered and mediate early antiviral responses. Type II interferons (interferon ) activate macrophages, interact with cells of the adaptive immune system and support the generation of TH1 cells. The TNF family of cytokines act as trimers and may occur in either soluble or membrane-bound forms. FasL, a TNF family member, interacts with its receptor, Fas, to stimulate apoptosis in the recipient cell. This interaction is important at the close of the immune response. TNF interacts with the TNF-R1 receptor on the surface of the cell to induce either apoptosis or survival, depending on the physiological environment. The IL-17 family of cytokines has been defined quite recently, and its members are primarily proinflammatory in action. Chemokines act on GPCR-coupled receptors to promote chemoattraction, the movement of immune system cells into, within, and out of lymphoid organs. Naturally occurring and pathogen-derived inhibitors of cytokine function may modulate their activity in vivo. Levels of inflammatory cytokines such as IL-1, IL-6, and TNF may be increased in certain disease states such as rheumatoid arthritis, and such diseases are susceptible to treatment with drugs that inhibit cytokine activities. R E F E R E N C E S Abram, C. L., and C. A. Lowell. 2009. The ins and outs of leukocyte integrin signaling. Annual Review of Immunology 27:339–362. Alcami, A. 2003. Structural basis of the herpesvirus M3-chemokine interaction. Trends in Microbiology 11:191–192. Ahmed, M., and S. L. Gaffen. 2010. IL-17 in obesity and adipogenesis. Cytokine & Growth Factor Reviews 21:449–453. Arend, W.P., G. Palmer, and C. Gabay. 2008. IL-1, IL-18, and IL-33 families of cytokines. Immunological Reviews 223:20–38. Receptors and Signaling: Cytokines and Chemokines | CHAPTER 4 139 Botto, S., et al. 2011. IL-6 in human cytomegalovirus secretome promotes angiogenesis and survival of endothelial cells through the stimulation of survivin. Blood 117:352–361. Micheau, M., and J. Tschopp. 2003. Induction of TNF Receptor 1-mediated apoptosis via two sequential signaling complexes. Cell 114:181–190. Boulanger, M. J., and K. C. Garcia. 2004. Shared cytokine signaling receptors: Structural insights from the gp130 system. Advances in Protein Chemistry 68:107–146. Raab, M., et al. 2010. T cell receptor “inside-out” pathway via signaling module SKAP1-RapL regulates T cell motility and interactions in lymph nodes. Immunity 32:541–556. Crabtree, G. R., S. Gillis, K. A. Smith, and A. Munck. 1980. Mechanisms of glucocorticoid-induced immunosuppression: Inhibitory effects on expression of Fc receptors and production of T-cell growth factor. Journal of Steroid Biochemistry 12:445–449. Raman, D., T. Sobolik-Delmaire, and A. Richmond. 2011. Chemokines in health and disease. Experimental Cell Research 317:5755–89. Crispin, J. C., and G. C. Tsokos. 2009. Transcriptional regulation of IL-2 in health and autoimmunity. Autoimmunity Reviews 8:190–195. Dandona, P., A. Aljada, and A. Bandyopadhyay. 2004. Inflammation: The link between insulin resistance, obesity and diabetes. Trends in Immunology 25:4–7. Dandona, P., A. Aljada, A. Chaudhuri, P. Mohanty, and G. Rajesh. 2004. A novel view of metabolic syndrome. Metabolic Syndrome and Related Disorders 2:2–8. Eisenbarth, S. C., and R. A. Flavell. 2009. Innate instruction of adaptive immunity revisited: The inflammasome. EMBO Molecular Medicine 1:92–98. Fickenscher, H., et al. 2002. The interleukin-10 family of cytokines. Trends in Immunology 23:89–96. Gaffen, S. L. (2009). Structure and signalling in the IL-17 receptor family. Nature Reviews. Immunology 9:556–567. Gee, K., C. Guzzo, N. F. Che Mat, W. Ma, and A. Kumar. 2009. The IL-12 family of cytokines in infection, inflammation and autoimmune disorders. Inflammation & Allergy Drug Targets 8:40–52. Gillis, S., R. Mertelsmann, and M. A. Moore. 1981. T-cell growth factor (interleukin 2) control of T-lymphocyte proliferation: Possible involvement in leukemogenesis. Transplantation Proceedings 13:1884–1890. Guilherme, A., J. V. Virbasius, V. Puri, and M. P. Czech. 2008. Adipocyte dysfunctions linking obesity to insulin resistance and type 2 diabetes. Nature Reviews. Molecular Cell Biology 9:367–377. Horng, T., and G. S. Hotamisligil. 2011. Linking the inflammasome to obesity-related disease. Nature Medicine 17:164–165. Hotamisligil, G. S. 2003. Inflammatory pathways and insulin action. International Journal of Obesity and Related Metabolic Disorders 27:Suppl 3, S53–55. Hotamisligil, G. S., N. S. Shargill, and B. M. Spiegelman. 1993. Adipose expression of tumor necrosis factor-alpha: Direct role in obesity-linked insulin resistance. Science 259:87–91. Rochman, Y., R. Spolski, and W. J. Leonard. 2009. New insights into the regulation of T cells by gamma(c) family cytokines. Nature Reviews. Immunology 9:480–490. Rot, A., and U. H. von Andrian. 2004. Chemokines in innate and adaptive host defense: Basic chemokinese grammar for immune cells. Annual Review of Immunollogy 22:891–928. Sadler, A. J., and B. R. Williams. 2008. Interferon-inducible antiviral effectors. Nature Reviews. Immunology 8:559–568. Skalnikova, H., J. Motlik, S. J. Gadher, and H. Kovarova. Mapping of the secretome of primary isolates of mammalian cells, stem cells and derived cell lines. Proteomics 11:691–708. Skiniotis, G., P. J. Lupardus, M. Martick, T. Walz, T., and K. C. Garcia. 2008. Structural organization of a full-length gp130/ LIF-R cytokine receptor transmembrane complex. Molecular Cell 31:737–748. Smith, K. A., P. E. Baker, S. Gillis, and F. W. Ruscetti. (1980). Functional and molecular characteristics of T-cell growth factor. Molecular Immunology 17:579–589. Stull, D., and S. Gillis. 1981. Constitutive production of interleukin 2 activity by a T cell hybridoma. Journal of Immunology 126:1680–1683. Walczak, H. 2011. TNF and ubiquitin at the crossroads of gene activation, cell death, inflammation and cancer. Immunological Reviews 244:9–28. Wellen, K. E., and G. S. Hotamisligil. 2003. Obesity-induced inflammatory changes in adipose tissue. The Journal of Clinical Investigation 112:1785–1788. Useful Web Sites www.invitrogen.com www.miltenyibiotec.com/cytokines www.prospecbio.com/Cytokines www.peprotech.com www.rndsystems.com Many companies that sell recom- Jones, L. L., and D. A. Vignali. 2011. Molecular interactions within the IL-6/IL-12 cytokine/receptor superfamily. Immunologic Research 51:5–14. binant cytokines or cytokine-related products provide useful information on their websites, or in print copy. The preceding are a few that are particularly helpful. Li, W. X. 2008. Canonical and non-canonical JAK-STAT signaling. Trends in Cell Biology 18:545–551. www.jakpathways.com/understandingjakpathways A useful animation of JAK-STAT signaling. 140 PA R T I | Introduction www.youtube.com/watch?vZUUfdP87Ssg&featu rerelated A rather wonderful movie of chemotaxis. You Tube. This one shows a macrophage recognizing a pathogen and releasing cytokines in response. www.youtube.com/watch?vEpC6G_DGqkI& featurerelated A striking movie of a neutrophil chas- www.netpath.org A curated set of pathways, with infor- ing a bacterium. mation on interacting proteins. Many interleukin pathways are included. www.youtube.com/watch?vKiLJl3NwmpU An increasing number of medical animations are available on S T U D Y Q U E S T I O N S 1. Distinguish between a hormone, a cytokine, a chemokine, and a growth factor. What functional attributes do they share, and what properties can be used to discriminate among them? 2. Measurement of the blood concentration of a particular cytokine reveals that it is rarely present above 1010 M, even under the conditions of an ongoing immune response. However, when you measure the affinity of the cognate receptor, you discover that its dissociation constant is close to 108 M, implying that the receptor occupancy must rarely exceed 1%. How do you account for this discrepancy? 3. Describe how dimerization and phosphorylation of intra- cellular signaling molecules contribute to activation of cells by Type 1 cytokines. 4. Define the terms pleiotropy, synergy, redundancy, antago- nism, and cascade induction as they apply to cytokine action. system ensure that only T cells that have been stimulated by antigen are susceptible to IL-2 signaling? 10b. The following diagram represents the results of a flow cytometry experiment in which mouse spleen cells were stained with antibodies directed against different components of the IL-2R. The more antibody that binds to the cells, the further they move along the relevant axis. The number of cells stained with fluorescein-conjugated anti IL-2R antibodies are shown along the x-axis of the flow cytometry plot, and cells that stain with phycoerythrinlabeled antibodies to the  subunit of the IL-2 receptor move along the Y-axis. We have drawn for your reference a circle that represents cells that stain with neither antibody. On this plot, draw, as circles, and label where you would expect to find the populations representing unstimulated T cells and T cells after antigen activation, after treatment with the two fluorescent labels described above. tion of the location of a lymphocyte? 6. Cytokines signaling through the Class I cytokine receptors can compete with one another, even though the recognition units of the receptors are different. Explain. 7. Describe one mechanism by which Type I interferons “interfere” with the production of new viral particles. 8. Signaling by tumor necrosis factor can paradoxically lead to cell activation or cell death. Explain how, by drawing diagrams of the relevant signaling pathways. 9. Describe two examples of ways in which vertebrates tune Populations staining with anti-IL-2Rα antibodies labeled with phycoerythrin 5. How might receipt of a cytokine signal result in the altera- down the intensity of their own cytokine signaling. 10a. The cytokine IL-2 is capable of activating all T cells to proliferation and differentiation. How does the immune Populations staining with anti-IL-2Rβ and anti-IL-2Rγ antibodies labeled with fluorescein 5 Innate Immunity V ertebrates are protected by both innate immunity and adaptive immunity. In contrast to adaptive immune responses, which take days to arise following exposure to antigens, innate immunity consists of the defenses against infection that are ready for immediate action when a host is attacked by a pathogen (viruses, bacteria, fungi, or parasites). The innate immune system includes anatomical barriers against infection—both physical and chemical—as well as cellular responses (Overview Figure 5-1). The main physical barriers—the body’s first line of defense—are the epithelial layers of the skin and of the mucosal and glandular tissue surfaces connected to the body’s openings; these epithelial barriers prevent infection by blocking pathogens from entering the body. Chemical barriers at these surfaces include specialized soluble substances that possess antimicrobial activity as well as acid pH. Pathogens that breach the physical and chemical barriers due to damage to or direct infection of the epithelial cell layer can survive in the extracellular spaces (some bacteria, fungi, and parasites) or they can infect cells (viruses and some bacteria and parasites), eventually replicating and possibly spreading to other parts of the body. The cellular innate immune responses to invasion by an infectious agent that overcomes the initial epithelial barriers are rapid, typically beginning within minutes of invasion. These responses are triggered by cell surface or intracellular receptors that recognize conserved molecular components of pathogens. Some white blood cell types (macrophages and neutrophils) are activated to rapidly engulf and destroy extracellular microbes through the process of phagocytosis. Other receptors induce the production of proteins and other substances that have a variety of beneficial effects, including direct antimicrobial activity and the recruitment of fluid, cells, and molecules to sites of infection. This influx causes swelling and other physiological changes that collectively are called inflammation. Such local innate and inflammatory responses usually are beneficial for eliminating pathogens and damaged or dead cells and promoting healing. For example, increased levels of antimicrobial substances and A macrophage (yellow) on the surface of a blood vessel (red) binds and phagocytoses bacteria (orange). [Dennis Kunkel Microscopy, Inc./Visuals Unlimited, Inc.] ■ Anatomical Barriers to Infection ■ Phagocytosis ■ Induced Cellular Innate Responses ■ Inflammatory Responses ■ Natural Killer Cells ■ ■ ■ Regulation and Evasion of Innate and Inflammatory Responses Interactions Between the Innate and Adaptive Immune Systems Ubiquity of Innate Immunity phagocytic cells help to eliminate the pathogens, and dendritic cells take up pathogens for presentation to lymphocytes, activating adaptive immune responses. Natural killer cells recruited to the site can recognize and kill virus-infected, altered, or stressed cells. However, in some situations these innate and inflammatory responses can be harmful, leading to local or systemic consequences that can cause tissue damage and occasionally death. To prevent these potentially harmful responses, regulatory mechanisms have evolved that usually limit such adverse effects. Despite the multiple layers of the innate immune system, some pathogens may evade the innate defenses. On call in vertebrates is the adaptive immune system, which counters infection with a specific tailor-made 141 142 PA R T I I | Innate Immunity 5-1 OVERVIEW FIGURE Innate Immunity Physical barriers to infection Epithelial layers of skin and mucosal/glandular tissues H+ H+ Chemical barriers to infection Pathogens H+ H+ Acidic pH and anti-microbial proteins and peptides H+ + H + H H+ Damage/infection Cellular responses to infection Pathogens Macrophage Infected cell Phagocytosis and degradation Dendritic cell Killing by NK cell Binding to cells Activation of adaptive immune responses NK cell Anti-microbial substances (e.g., peptides, interferons) Pathogen elimination Cell activation Inflammation: recruitment and activation of protective cells and molecules (e.g., complement) to the infection site Key elements of innate immunity include the physical and chemical barriers that prevent infection, provided by the epithelial cell layers of the skin, mucosal tissues (e.g., gastrointestinal, respiratory, and urogenital tracts), and glandular tissues (e.g., salivary, lacrimal, and mammary glands). Once pathogens enter the body, such as through a breach in an epithelial layer, they are confronted by an array of cells with cell surface and intracellular receptors that recognize pathogen components and trigger a variety of cellular responses. Pathogen recognition by these receptors activates some cells to phagocytose and degrade the pathogen, and many cells response to the attacking pathogen. This attack occurs in the form of B and T lymphocytes, which generate antibodies, and effector T cells that specifically recognize and neutralize or eliminate the invaders. Table 5-1 compares innate and adaptive immunity. While innate Antibodies T-cell responses Cytokines and chemokines Systemic effects (e.g., fever) are activated through their receptors to produce a variety of antimicrobial substances that kill pathogens, as well as cytokine and chemokine proteins that recruit cells, molecules, and fluid to the site of infection, leading to swelling and other symptoms collectively known as inflammation. The innate natural killer (NK) cells recognize and kill some virus-infected cells. Cytokines and chemokines can cause systemic effects that help to eliminate an infection, and also contribute—along with dendritic cells that carry and present pathogens to lymphocytes—to the activation of adaptive immune responses. immunity is the most ancient form of defense, found in all multicellular plants and animals, adaptive immunity is a much more recent evolutionary invention, having arisen in vertebrates. In these animals, adaptive immunity complements a well-developed system of innate immune Innate Immunity TABLE 5-1 | CHAPTER 5 143 Innate and adaptive immunity Attribute Innate immunity Adaptive immunity Response time Minutes/hours Days Specificity Specific for molecules and molecular patterns associated with pathogens and molecules produced by dead/damaged cells Highly specific; discriminates between even minor differences in molecular structure of microbial or nonmicrobial molecules Diversity A limited number of conserved, germ line– encoded receptors Highly diverse; a very large number of receptors arising from genetic recombination of receptor genes in each individual Memory responses Some (observed in invertebrate innate responses and mouse/human NK cells) Persistent memory, with faster response of greater magnitude on subsequent exposure Self/nonself discrimination Perfect; no microbe-specific self/nonself patterns in host Very good; occasional failures of discrimination result in autoimmune disease Soluble components of blood Many antimicrobial peptides, proteins, and other mediators Antibodies and cytokines Major cell types Phagocytes (monocytes, macrophages, neutrophils), natural killer (NK) cells, other leukocytes, epithelial and endothelial cells T cells, B cells, antigen-presenting cells mechanisms that share important features with those of our invertebrate ancestors. A large and growing body of research has revealed that as innate and adaptive immunity have co-evolved in vertebrates, a high degree of interaction and interdependence has arisen between the two systems. Recognition by the innate immune system not only kicks off the adaptive immune response but also helps to ensure that the type of adaptive response generated will be effective for the invading pathogen. This chapter describes the components of the innate immune system—physical barriers, chemical agents, and a battery of protective cellular responses carried out by numerous cell types—and illustrates how they act together to defend against infection. We conclude with an overview of innate immunity in plants and invertebrates. Anatomical Barriers to Infection The most obvious components of innate immunity are the external barriers to microbial invasion: the epithelial layers that insulate the body’s interior from the pathogens of the exterior world. These epithelial barriers include the skin and the tissue surfaces connected to the body’s openings: the mucous epithelial layers that line the respiratory, gastrointestinal, and urogenital tracts and the ducts of secretory glands such as the salivary, lacrimal, and mammary glands (which produce saliva, tears, and milk, respectively) (Figure 5-2). Skin and other epithelia provide a kind of living “plastic wrap” that encases and protects the inner domains of the body from infection. But these anatomical barriers are more than just passive wrappers. They contribute to physical and mechanical processes that help the body shed pathogens and also generate active chemical and biochemical defenses by synthesizing and deploying molecules, including peptides and proteins, that have or induce antimicrobial activity. Epithelial Barriers Prevent Pathogen Entry into the Body’s Interior The skin, the outermost physical barrier, consists of two distinct layers: a thin outer layer, the epidermis, and a thicker layer, the dermis. The epidermis contains several tiers of tightly packed epithelial cells; its outer layer consists of mostly dead cells filled with a waterproofing protein called keratin. The dermis is composed of connective tissue and contains blood vessels, hair follicles, sebaceous glands, sweat glands, and scattered myeloid leukocytes such as dendritic cells, macrophages, and mast cells. In place of skin, the respiratory, gastrointestinal, and urogenital tracts and the ducts of the salivary, lacrimal, and mammary glands are lined by strong barrier layers of epithelial cells stitched together by tight junctions that prevent pathogens from squeezing between them to enter the body. A number of nonspecific physical and chemical defense mechanisms also contribute to preventing the entry of pathogens through the epithelia in these secretory tissues. For example, the secretions of these tissues (mucus, urine, saliva, tears, and milk) wash away potential invaders and also contain antibacterial and antiviral substances. Mucus, the viscous fluid secreted by specialized cells of the mucosal epithelial layers, entraps foreign microorganisms. In the 144 PA R T I I | Innate Immunity Organ or tissue Innate mechanisms protecting skin/epithelium Skin Antimicrobial peptides, fatty acids in sebum Mouth and upper alimentary canal Enzymes, antimicrobial peptides, and sweeping of surface by directional flow of fluid toward stomach Stomach Low pH, digestive enzymes, antimicrobial peptides, fluid flow toward intestine Small intestine Digestive enzymes, antimicrobial peptides, fluid flow to large intestine Large intestine Normal intestinal flora compete with invading microbes, fluid/feces expelled from rectum Airway and lungs Cilia sweep mucus outward, coughing, sneezing expel mucus, macrophages in alveoli of lungs Urogenital tract Flushing by urine, aggregation by urinary mucins; low pH, anti-microbial peptides, proteins in vaginal secretions Salivary, lacrimal, and mammary glands Flushing by secretions; anti-microbial peptides and proteins in vaginal secretions Skin Lacrimal glands Mouth Salivary glands Airway Lung Epithelial lining of airway and lung Epithelial lining of alimentary canal Mammary glands Stomach Large intestine Small intestine Rectum Urogenital tract In addition to serving as physical barriers, the skin and mucosal and glandular epithelial layers are defended against microbial coloniza- tion by a variety of mechanisms: mechanical (cilia, fluid flow, smooth muscle contraction), chemical (pH, enzymes, antimicrobial peptides), and cellular (resident macrophages and dendritic cells). lower respiratory tract, cilia, hairlike protrusions of the cell membrane, cover the epithelial cells. The synchronous movement of cilia propels mucus-entrapped microorganisms from these tracts. Coughing is a mechanical response that helps us get rid of excess mucus, with trapped microorganisms, that occurs in many respiratory infections. The flow of urine sweeps many bacteria from the urinary tract. With every meal, we ingest huge numbers of microorganisms, but they must run a gauntlet of defenses in the gastrointestinal tract that begins with the antimicrobial compounds in saliva and in the epithelia of the mouth and includes the hostile mix of digestive enzymes and acid found in the stomach. Similarly, the acidic pH of vaginal secretions is important in providing protection against bacterial and fungal pathogens. Some organisms have evolved ways to evade these defenses of the epithelial barriers. For example, influenza FIGURE 5-2 Skin and other epithelial barriers to infection. Innate Immunity FIGURE 5-3 Electron micrograph of E. coli bacteria adhering to the surface of epithelial cells of the urinary tract. E. coli is an intestinal bacterial species that causes urinary tract infections affecting the bladder and kidneys. [From N. Sharon and H. Lis, 1993, Scientific American 268(1):85; courtesy of K. Fujita; Matthew A. Mulvey, Joel D. Schilling, Juan J. Martinez, and Scott J. Hultgren, Bad bugs and beleaguered bladders: Interplay between uropathogenic Escherichia coli and innate host defenses. PNAS August 1, 2000 vol. 97 no. 16 8829-8835. Copyright 2000 National Academy of Sciences, U.S.A.] virus has a surface molecule that enables it to attach firmly to cells in mucous membranes of the respiratory tract, preventing the virus from being swept out by the ciliated epithelial cells. Neisseria gonorrhoeae, the bacteria that causes gonorrhea, binds to epithelial cells in the mucous membrane of the urogenital tract. Adherence of these and other bacteria to mucous membranes is generally mediated by hairlike protrusions on the bacteria called fimbriae or pili that have evolved the ability to bind to certain glycoproteins or glycolipids only expressed by epithelial cells of the mucous membrane of particular tissues (Figure 5-3). | CHAPTER 5 145 (which also incudes calprotectin) with potent antibacterial activity against Escherichia coli, an enteric (intestinal) bacterial species. This finding answered a long-standing question: Why is human skin resistant to colonization by E. coli despite exposure to it from fecal matter resulting from lack of cleanliness or poor sanitation? As shown in Figure 5-4, incubation of E. coli on human skin for as little as 30 minutes kills the bacteria. Antimicrobial proteins can show some specificity toward particular pathogens. For example, psoriasin is a key antimicrobial for E. coli on the skin (and also on the tongue), but, as shown in Figure 5-4, it does not kill Staphylococcus aureus (a major cause of food poisoning and skin infections), whereas the related protein calprotectin kills S. aureus but not E. coli. S. aureus E. coli Inoculation 30 minutes Fresh culture plates Incubate Antimicrobial Proteins and Peptides Kill Would-be Invaders To provide strong defense at these barrier layers, epithelial cells secrete a broad spectrum of proteins and peptides that provide protection against pathogens. The capacity of skin and other epithelia to produce a wide variety of antimicrobial agents on an ongoing basis is important for controlling the microbial populations on these surfaces, as breaks in these physical barriers from wounds provide routes of infection that would be readily exploited by pathogenic microbes if not defended by biochemical means. Among the antimicrobial proteins produced by the skin and other epithelia in humans (Table 5-2), several are enzymes and binding proteins that kill or inhibit growth of bacterial and fungal cells. Lysozyme is an enzyme found in saliva, tears, and fluids of the respiratory tract that cleaves the peptidoglycan components of bacterial cell walls. Lactoferrin and calprotectin are two proteins that bind and sequester metal ions needed by bacteria and fungi, limiting their growth. Among the many antibacterial agents produced by human skin, recent research has identified psoriasin, a small protein of the S-100 family FIGURE 5-4 Psoriasin prevents colonization of the skin by E. coli. Skin secretes psoriasin, an antimicrobial protein that kills E. coli. Fingertips of a healthy human were inoculated with Staphylococcus aureus (S. aureus) and E. coli. After 30 minutes, the fingertips were pressed on a nutrient agar plate and the number of colonies of S. aureus and E. coli determined. Almost all of the inoculated E. coli were killed; most of the S. aureus survived. [Photograph courtesy of Nature Immunology; from R. Gläser et al., 2005, Nature Immunology 6:57–64.] 146 TABLE 5-2 PA R T I I | Innate Immunity Some human antimicrobial proteins and peptides at epithelial surfaces Proteins and peptides* Location Antimicrobial activities Lysozyme Mucosal/glandular secretions (e.g., tears, saliva, respiratory tract) Cleaves glycosidic bonds of peptidoglycans in cell walls of bacteria, leading to lysis Lactoferrin Mucosal/glandular secretions (e.g., milk, intestine mucus, nasal/respiratory and urogenital tracts) Binds and sequesters iron, limiting growth of bacteria and fungi; disrupts microbial membranes; limits infectivity of some viruses Secretory leukocyte protease inhibitor Skin, mucosal/glandular secretions (e.g., intestines, respiratory, and urogenital tracts, milk) Blocks epithelial infection by bacteria, fungi, viruses; antimicrobial S100 proteins, e.g.: - psoriasin - calprotectin Skin, mucosal/glandular secretions (e.g., tears, saliva/tongue, intestine, nasal/ respiratory and urogenital tracts) Defensins ( and ) Skin, mucosal epithelia (e.g., mouth, intestine, nasal/respiratory tract, urogenital tract) Disrupt membranes of bacteria, fungi, protozoan parasites, and viruses; additional toxic effects intracellularly; kill cells and disable viruses Cathelicidin (LL37)** Mucosal epithelia (e.g., respiratory tract, urogenital tract) Disrupts membranes of bacteria; additional toxic effects intracellularly; kills cells. Surfactant proteins SP-A, SP-D Secretions of respiratory tract, other mucosal epithelia Block bacterial surface components; promotes phagocytosis - Disrupts membranes, killing cells - Binds and sequesters divalent cations (e.g., manganese and zinc), limiting growth of bacteria and fungi *Examples listed in this table are all produced by cells in the epithelia of mucosal and glandular tissues; examples of prominent epithelial sites are listed. Most proteins and peptides are produced constitutively at these sites, but their production can also be increased by microbial or inflammatory stimuli. Many are also produced constitutively in neutrophils and stored in granules. In addition, synthesis and secretion of many of these molecules may be induced by microbial components during innate immune responses by various myeloid leukocyte populations (monocytes, macrophages, dendritic cells, and mast cells). **While some mammals have multiple cathelicidins, humans have only one. Another major class of antimicrobial components secreted by skin and other epithelial layers is comprised of antimicrobial peptides, generally less than 100 amino acids long, which are an ancient form of innate immunity present in vertebrates, invertebrates, plants, and even some fungi. The discovery that vertebrate skin produces antimicrobial proteins came from studies in frogs, where glands in the skin were shown to secrete peptides called magainins that have potent antimicrobial activity against bacteria, yeast, and protozoans. Antimicrobial peptides generally are cysteine-rich, cationic, and amphipathic (containing both hydrophilic and hydrophobic regions). Because of their positive charge and amphipathic nature, they interact with acidic phospholipids in lipid bilayers, disrupting membranes of bacteria, fungi, parasites, and viruses. They then can enter the microbes, where they have other toxic effects, such as inhibiting the synthesis of DNA, RNA, or proteins, and activating antimicrobial enzymes, resulting in cell death. The main types of antimicrobial peptides found in humans are - and -defensins and cathelicidin. Human defensins kill a wide variety of bacteria, including E. coli, S. aureus, Streptococcus pneumoniae, Pseudomonas aeruginosa, and Hemophilus influenzae. Antimicrobial peptides also attack the lipoprotein envelope of enveloped viruses such as influ- enza virus and some herpesviruses. Defensins and cathelicidin LL-37 (the only cathelicidin expressed in humans) are secreted constitutively by epithelial cells in many tissues, as well as stored in neutrophil granules where they contribute to killing phagocytosed microbes. Recent studies have shown that human -defensin antimicrobial peptides secreted into the gut by intestinal epithelial Paneth cells, located in the deep valleys (crypts) between the villi, are important for maintaining beneficial bacterial flora that are necessary for normal intestinal immune system functions. As we will see later, production of these antimicrobial peptides also can be induced in many epithelial and other cell types by the binding of microbial components to cellular receptors. The epithelium of the respiratory tract secretes a variety of lubricating lipids and proteins called surfactants. Two surfactant proteins, SP-A and SP-D, which are present in the lungs as well as in the secretions of some other mucosal epithelia, are members of a class of microbe-binding proteins called collectins. SP-A and SP-D bind differentially to sets of carbohydrate, lipid, and protein components of microbial surfaces and help to prevent infection by blocking and modifying surface components and promoting pathogen clearance. For example, they differentially bind two alternating states of the lung pathogen Klebsiella pneumonia that differ in whether or not Innate Immunity they are coated with a thick polysaccharide capsule: SP-A binds the complex polysaccharides coating many of the capsulated forms, while SP-D only binds the exposed cell wall lipopolysaccharide of the nonencapsulated form. Phagocytosis Despite the strong defenses of our protective epithelial layers, some pathogens have evolved strategies to penetrate these defenses, and epithelia may be disrupted by wounds, abrasions, and insect bites that may transmit pathogens. Once pathogens penetrate through the epithelial barrier layers into the tissue spaces of the body, an array of cellular membrane receptors and soluble proteins that recognize microbial components play the essential roles of detecting the pathogen and triggering effective defenses against it. Phagocytic cells make up the next line of defense against pathogens that have penetrated the epithelial cell barriers. Macrophages, neutrophils, and dendritic cells in tissues and monocytes in the blood are the main cell types that carry out phagocytosis—the cellular uptake (eating) of particulate materials such as bacteria—a key mechanism for eliminating pathogens. This major role of the cells attracted to the site of invading organisms is evolutionarily ancient, present in invertebrates as well as vertebrates. Elie Metchnikoff initially described the process of phagocytosis in the 1880s using cells from starfish (echinoderm invertebrates) similar to vertebrate white blood cells and ascribed to phagocytosis a major role in immunity. He was correct in this conclusion; we now know that defects in phagocytosis lead to severe immunodeficiency. As described in Chapter 2, most tissues contain resident populations of macrophages that function as sentinels for | CHAPTER 5 147 the innate immune system. Through various cell surface receptors they recognize microbes such as bacteria, extend their plasma membrane to engulf them, and internalize them in phagosomes (endosomes resulting from phagocytosis, Figure 5-5). Lysosomes then fuse with the phagosomes, delivering agents that kill and degrade the microbes. Neutrophils are a second major type of phagocyte, usually recruited to sites of infection. Finally, dendritic cells also can bind and phagocytose microbes. As will be described later in this chapter and more extensively in Chapters 8 and 11, uptake and degradation of microbes by dendritic cells play key roles in the initiation of adaptive immune responses. In addition to triggering phagocytosis, various receptors on phagocytes recognize microbes and activate the production of a variety of molecules that contribute in other ways to eliminating infection, as will be described later in this chapter. A phagocyte’s recognition of microbes and the responses that result are shown in Overview Figure 5-6. Microbes are Recognized by Receptors on Phagocytic Cells How does a phagocytic cell recognize microbes, triggering their phagocytosis? Phagocytes express on their surfaces a variety of receptors, some of which directly recognize specific conserved molecular components on the surfaces of microbes, such as cell wall components of bacteria and fungi. These conserved motifs, usually present in many copies on the surface of a bacterium, fungal cell, parasite, or virus particle, are called pathogen-associated molecular patterns (PAMPs). Note that they can be expressed by microbes whether or not the microbes are pathogenic (cause (b) (a) 1 2 Bacterium becomes attached to membrane evaginations called pseudopodia. Bacterium is ingested, forming phagosome. 3 Phagosome fuses with lysosome. 4 Bacterium is killed and then digested by lysosomal enzymes. 5 Digestion products are released from cell. FIGURE 5-5 Phagocytosis. (a) Scanning electron micrograph of alveolar macrophage phagocytosis of E. coli bacteria on the outer surface of a blood vessel in the lung pleural cavity. (b) Steps in the phagocytosis of a bacterium. [Part (a) from Dennis Kunkel Microscopy, Inc./Visuals Unlimited, Inc.] 5-6 OVERVIEW FIGURE Effectors of Innate Immune Responses to Infection Phagocytes (neutrophils, dendritic cells, macrophages) Pathogen-associated molecular patterns (PAMPs) PAMPs recognized by pattern recognition receptors (PRRs) Pathogen killed and degraded in lysosomes Phagocyte activated to produce anti-microbial components, followed by secretion of inflammationpromoting cytokines and chemokines Phagocytosis Pathogen C-reactive protein (CRP) Complement proteins SP-A, SP-D, Mannosebinding lectin (MBL) Opsonins are recognized by opsonin receptors, enhancing phagocytosis Innate initiation of adaptive response Dendritic cell PRRs recognize PAMPs, activating phagocytosis and signaling pathways. Dendritic cells migrate to lymph nodes, carrying intact or degraded pathogens. Antigen fragments bound to cell surface MHC proteins are recognized by T cells. T cell Activated T cells initiate adaptive responses. Microbial invasion brings many effectors of innate immunity into play. Entry of microbial invaders through lesions in epithelial barriers generates inflammatory signals and exposes the invaders to attack by different effector molecules and cells. Microbes with surface components recognized by C-reactive protein (CRP), mannosebinding lectin (MBL), or surfactant proteins A or D (SP-A and SP-D) are bound by these opsonizing molecules, marking the microbes for phagocytosis by neutrophils and macrophages. Some bacteria and fungi can activate complement directly, or via bound CRP or MBL, leading to further opsonization or direct lysis. Inflammatory signals cause phagocytes such as monocytes and neutrophils to bind to the walls of blood vessels, extravasate, and move to the site of infection, where they phagocytose and kill infecting Antimicrobial peptides Membrane damage kills pathogen Opsonized pathogen CRP, MBL, complement proteins activate complement pathway Complement destroys membrane, stimulates inflammation, attracts neutrophils and other cells microorganisms. Binding of microbes to receptors on phagocytes activates phagocytosis and production of additional antimicrobial and proinflammatory molecules that intensify the response, in part by recruiting more phagocytes and soluble mediators (CRP, MBL, and complement) from the bloodstream to the site of infection. Inset: Dendritic cells bind microbes via receptors and are activated to mature; they also internalize and degrade microbes. These dendritic cells migrate through lymphatic vessels to nearby lymph nodes, where they present antigen-derived peptides on their MHC proteins to T cells. Antigen-activated T cells then initiate adaptive immune responses against the pathogen. Cytokines produced during innate immune responses also support and direct the adaptive immune responses to infection. Innate Immunity TABLE 5-3 | CHAPTER 5 149 Human receptors that trigger phagocytosis Receptor type on phagocytes Examples Ligands Pattern recognition receptors C-type lectin receptors (CLRs) Scavenger receptors Microbial ligands (found on microbes) Mannose receptor Mannans (bacteria, fungi, parasites) Dectin 1 -glucans (fungi, some bacteria) DC-SIGN Mannans (bacteria, fungi, parasites) SR-A Lipopolysaccharide (LPS), lipoteichoic acid (LTA) (bacteria) SR-B LTA, lipopeptides, diacylglycerides (bacteria), -glucans (fungi) Microbe-binding opsonins (soluble; bind to microbes) Opsonin receptors Collagen-domain receptor CD91/calreticulin Collectins SP-A, SP-D, MBL; L-ficolin; C1q Complement receptors CR1, CR3, CR4, CRIg, C1qRp Complement components and fragments* Immunoglobulin Fc receptors FcR Specific IgA antibodies bound to antigen# FcRs Specific IgG antibodies bound to antigen;# C-reactive protein * See Table 6-3 for specific complement components or fragments that are bound by individual receptors # Opsonization of antibody-bound antigens is an adaptive immune response clearance mechanism disease); hence some researchers have started to use the more general term microbe-associated molecular patterns (MAMPs). The receptors that recognize PAMPs are called pattern recognition receptors (PRRs). Some PRRs that bind microbes and trigger phagocytosis of the bound microbes are listed at the top of Table 5-3, along with the PAMPs they recognize. As we will see later, there are other PRRs that, after PAMP binding, do not activate phagocytosis but trigger other types of responses. Most PAMPs that induce phagocytosis are cell wall components, including complex carbohydrates such as mannans and -glucans, lipopolysaccharides (LPS), other lipid-containing molecules, peptidoglycans, and surface proteins. As shown in Figure 5-6, activation of phagocytosis can also occur indirectly, by phagocyte recognition of soluble proteins that have bound to microbial surfaces, thus enhancing phagocytosis, a process called opsonization (from the Greek word for “to make tasty”). Many of these soluble phagocytosis-enhancing proteins (called opsonins) also bind to conserved, repeating components on the surfaces of microbes such as carbohydrate structures, lipopolysaccharides, and viral proteins; hence they are sometimes referred to as soluble pattern-recognition proteins. Once bound to microbe surfaces, opsonins are recognized by membrane opsonin receptors on phagocytes, activating phagocytosis (see Table 5-3, bottom). A variety of soluble proteins function as opsonins; many play other roles as well in innate immunity. For example, the two surfactant collectin proteins mentioned earlier, SP-A and SP-D, are found in the blood as well as in mucosal secretions in the lungs and elsewhere, where they function as opsonins. After binding to microbes they are recognized by the CD91 opsonin receptor (see Table 5-3) and promote phagocytosis by alveolar and other macrophage populations. This function of SP-A and SP-D contributes to clearance of the fungal respiratory pathogen Pneumocystis carinii, a major cause of pneumonia in individuals with AIDS. Mannose-binding lectin (MBL), a third collectin with opsonizing activity, is found in the blood and respiratory fluids. L-ficolin, a member of the ficolin family that is related to MBL and other collectins, is found in the blood, where it binds to acetylated sugars on microbes, including some streptococcal bacteria. The complement component C1q also functions as an opsonin, binding bacterial cell wall components such as lipopolysaccharides and some viral proteins. MBL (and other collectins), ficolins, and C1q share structural features, including similar polymeric structures with collagen-like shafts, but have recognition regions with different binding specificities (Figure 5-7). As a result of their structural similarities, all are bound by the CD91 opsonin receptor (see Table 5-3) and activate pathogen phagocytosis. Another opsonin, C-reactive protein (CRP), recognizes phosphorylcholine and carbohydrates on bacteria, fungi, and parasites, and is then bound by Fc receptors (FcRs) for IgG found on most phagocytes (see Chapter 3). Fc receptors also are important for the opsonizing activity of IgA antibodies and some IgG antibody subclasses. After binding specifically to antigens on microbe surfaces, the Fc regions of these antibodies can be recognized by specific FcRs, triggering phagocytosis. As an important mechanism by which the PA R T I I 150 | Innate Immunity (a) Mannose-binding lectin (c) C1 bound to LPS C1q Cysteine-rich region Collagen-like region MASP-2 C1r MASP-3 C1s Recognition domain Carbohydrate (b) H-ficolin LPS (d) C1 bound to lgG C1q MASP-2 MASP-3 Carbohydrate FIGURE 5-7 Structures of opsonins. (a) Mannose-binding lectin (MBL), a collectin, and (b) H-ficolin are polymers of multiple polypeptide chains, each containing an N-terminal cysteine-rich region followed by a collagen-like region, an -helical neck region, and a recognition domain. Both MBL and H-ficolin have bound MBL-associated serine proteases (MASPs), which become active after the recognition domains bind to specific carbohydrate residues on pathogen surfaces. The MASPs then activate the complement adaptive immune response clears antigens, opsonization by antibodies (not shown in Figure 5-6) will be discussed in Chapter 13. As mentioned above, among the most effective opsonins are several components of the complement system, which is described detail in Chapter 6. Present in both invertebrates and vertebrates, complement straddles both the innate and adaptive immune systems, indicating that it is ancient and important. In vertebrates, complement consists of more than 30 binding proteins and enzymes that function in a cascade of sequential activation steps. It can be triggered by several innate soluble pattern-recognition proteins (including the C1r C1s lgG pathway. C1, the first component of complement, has a multimeric structure similar to that of MBL and H-ficolin. C1q can bind LPS on bacterial cell walls (c) or antibodies bound to antigens (d). After C1 binds to the pathogen-associated LPS or antibodies, the associated C1r and C1s subunit proteases, which are similar to MASPs, become activated and can initiate complement activation. [Adapted from A. L. DeFranco, R. M. Locksley, and M. Robertson, 2007, Immunity (Primers in Biology), Sunderland, MA: Sinauer Associates.] first complement component, C1q, and the structurally related lectins MBL and ficolins, C-reactive protein, and properdin), as well as by microbe-bound antibodies generated by the adaptive immune response. As we will see in Chapter 6, phagocytosis is one of many important antimicrobial effects resulting from complement activation. The importance of MBL’s roles as both an opsonizer and an inducer of complement activation has been indicated by the effects of MBL deficiencies, which affect about 25% of the population. Individuals with MBL deficiencies are predisposed to severe respiratory tract infections, especially pneumococcal pneumonia. Interestingly, MBL deficiencies may | Innate Immunity be protective against tuberculosis, probably reflecting MBL’s opsonizing role in enhancing the phagocytosis of Mycobacterium tuberculosis, the route by which it infects macrophages, potentially leading to tuberculosis. Phagocytosed Microbes are Killed by Multiple Mechanisms The binding of microbes—bacteria, fungi, protozoan parasites, and viruses—to phagocytes via pattern recognition receptors or opsonins and opsonin receptors (see Figure 5-6) activates signaling pathways. These signaling pathways trigger actin polymerization, resulting in membrane extensions around the microbe particles and their internalization, forming phagosomes (see Figure 5-5). The phagosomes then fuse with lysosomes and, in neutrophils, with preformed primary and secondary granules (see Figure 2-2a). The resulting phagolysosomes contain an arsenal of antimicrobial agents that then kill and degrade the internalized microbes. These agents include antimicrobial proteins and peptides (including defensins and cathelicidins), low pH, acid-activated hydrolytic enzymes (including lysozyme and proteases), and specialized molecules that mediate oxidative attack. Oxidative attack on the phagocytosed microbes, which occurs in neutrophils, macrophages, and dendritic cells, employs highly toxic reactive oxygen species (ROS) and reactive nitrogen species (RNS), which damage intracellular components (Figure 5-8). The reactive oxygen species are generated by the phagocytes’ unique NADPH oxidase enzyme complex (also called phagosome oxidase), which is CHAPTER 5 151 activated when microbes bind to the phagocytic receptors. The oxygen consumed by phagocytes to support ROS production by NADPH oxidase is provided by a metabolic process known as the respiratory burst, during which oxygen uptake by the cell increases severalfold. NADPH oxidase converts oxygen to superoxide ion (•O2); other ROS generated by the action of additional enzymes are hydrogen peroxide (H2O2), and hypochlorous acid (HClO), the active component of household bleach. The generation of RNS requires the transcriptional activation of the gene for the enzyme inducible nitric oxide synthase (iNOS, or NOS2)—called that to distinguish it from related NO synthases in other tissues. Expression of iNOS is activated by microbial components binding to various PRRs. iNOS oxidizes L-arginine to yield L-citrulline and nitric oxide (NO), a potent antimicrobial agent. In combination with superoxide ion (•O2) generated by NADPH oxidase, NO produces an additional reactive nitrogen species, peroxynitrite (ONOO) and toxic S-nitrosothiols. Collectively the ROS and RNS are highly toxic to phagocytosed microbes due to the alteration of microbial molecules through oxidation, hydroxylation, chlorination, nitration, and S-nitrosylation, along with formation of sulfonic acids and destruction of iron-sulfur clusters in proteins. One example of how these oxidative species may be toxic to pathogens is the oxidation by ROS of cysteine sulfhydryls that are present in the active sites of many enzymes, inactivating the enzymes. ROS and RNS also can be released from activated neutrophils and macrophages and kill extracellular pathogens. Antimicrobial species generated from oxygen and nitrogen Reactive oxygen species (ROS) •O2 (superoxide anion) OH• (hydroxyl radical) H2O2 (hydrogen peroxide) HClO (hypochlorous acid) NADPH phagosome oxidase O2 Oxygen •O2 Superoxide anion Superoxide dismutase OH Hydroxyl radicals •OH H2O2 Hydrogen peroxide Myeloperoxidase HClO Hypochlorous acid Cl Chloride ion Reactive nitrogen species (RNS) NO (nitric oxide) NO2 (nitrogen dioxide) ONOO (peroxynitrite) Inducible nitric oxide synthase L-Arginine L-Citrulline NO Nitric oxide FIGURE 5-8 Generation of antimicrobial reactive oxygen and nitrogen species. In the cytoplasm of neutrophils, macrophages, and dendritic cells, several enzymes, including phagosome NADPH oxidase, transform molecular oxygen into highly reactive oxygen species (ROS) that have antimicrobial ONOO Peroxynitrite S-nitrosothiols NO2 Nitrogen dioxide activity. One of the products of this pathway, superoxide anion, can interact with a reactive nitrogen species (RNS), generated by inducible nitric oxide synthase (iNOS) to produce peroxynitrite, another RNS. NO can also undergo oxidation to generate the RNS nitrogen dioxide. 152 PA R T I I | Innate Immunity Evidence from genetic defects in humans and mice highlights the critical roles of these reactive chemical species in microbial elimination by phagocytic cells. The importance to antimicrobial defense of phagosomal NADPH oxidase and its products, ROS and RNS, is illustrated by chronic granulomatous disease (CGD). Patients afflicted with this disease have dramatically increased susceptibility to some fungal and bacterial infections, caused by defects in subunits of NADPH oxidase that destroy its ability to generate oxidizing species. In addition, studies with mice in which the genes encoding iNOS were knocked out have shown that nitric oxide and substances derived from it account for much of the antimicrobial activity of macrophages against bacteria, fungi, and parasites. These mice lost much of their usual ability to control infections by such intracellular pathogens as M. tuberculosis and Leishmania major, the intracellular protozoan parasite that causes leishmaniasis. Phagocytosis Contributes to Cell Turnover and the Clearance of Dead Cells Our discussion of phagocytosis thus far has focused on its essential roles in killing pathogens. As the body’s main scavenger cells, macrophages also utilize their phagocytic receptors to take up and clear cellular debris, cells that have died from damage or toxic stimuli (necrotic cell death) or from apoptosis (programmed cell death), and aging red blood cells. Considerable progress has been made in recent years in understanding the specific markers and receptors that trigger macrophage phagocytosis of dead, dying, and aging cells. Collectively the components of dead/dying cells and damaged tissues that are recognized by PRRs leading to their clearance are sometimes referred to as damage-associated molecular patterns (DAMPs). As the presence of these components may also be an indicator of conditions harmful to the body or may contribute to harmful consequences (such as autoimmune diseases), the D in DAMP also can refer to a “Danger” signal. Phagocytosis is the major mode of clearance of cells that have undergone apoptosis as part of developmental remodeling of tissues, normal cell turnover, or killing of pathogen-infected or tumor cells by innate or adaptive immune responses. Apoptotic cells attract phagocytes by releasing the lipid mediator lysophosphatidic acid, which functions as a chemoattractant. These dying cells facilitate their own phagocytosis by expressing on their surfaces an array of molecules not expressed on healthy cells, including phospholipids (such as phosphatidyl serine and lysophosphatidyl choline), proteins (annexin I), and altered carbohydrates. These DAMPs are recognized directly by phagocytic receptors such as the phosphatidyl serine receptor and scavenger receptor SR-A1. Other DAMPs are recognized by soluble pattern recognition molecules that function as opsonins, including the collectins MBL, SP-A, and SP-D mentioned earlier; various complement components; and the pentraxins C-reactive protein and serum amyloid protein. These opsonins are then recognized by opsonin receptors, activating phagocytosis and degradation of the apoptotic cells. An important additional activity of macrophages in the spleen and those in the liver (known as Kupffer cells) is to recognize, phagocytose, and degrade aging and damaged red blood cells. As these cells age, novel molecules that are recognized by phagocytes accumulate in their plasma membrane. Phosphatidyl serine flips from the inner to the outer leaflet of the lipid bilayer and is recognized by phosphatidyl serine receptors on phagocytes. Modifications of erythrocyte membrane proteins have also been detected that may promote phagocytosis. Obviously it is important for normal cells not to be phagocytosed, and accumulating evidence indicates that whether or not a cell is phagocytosed is controlled by sets of “eat me” signals—the altered membrane components (DAMPs) described above—and “don’t eat me” signals expressed by normal cells. Young, healthy erythrocytes avoid being phagocytosed by not expressing “eat me” signals, such as phosphatidyl serine, and also by expressing a “don’t eat me” signal, the protein CD47. CD47, expressed on many cell types throughout the body, is recognized by the SIRP- receptor on macrophages, which transmits signals that inhibit phagocytosis. Recent studies have shown that tumors use elevated CD47 expression to evade tumor surveillance and phagocytic elimination by the immune system. Increased expression of CD47 on all or most human cancers is correlated with tumor progression, probably because the CD47 activates SIRP-1-mediated inhibition of the phagocytosis of tumor cells by macrophages. This understanding of the role of CD47 in preventing phagocytosis is being used to develop novel therapies for certain cancers, such as using antibodies to block CD47 on tumor cells, which should then allow them to be phagocytosed and eliminated. Induced Cellular Innate Responses In addition to triggering their own uptake and killing by phagocytic cells, microbes induce a broad spectrum of cellular innate immune responses by a wide variety of cell types. Several families of pattern recognition receptors (PRRs) other than those described earlier as mediating phagocytosis (see Table 5-3) play major roles in innate immunity. As we will see in this section, these PRRs bind to PAMPs as well as to some endogenous (self) DAMPs and trigger signal-transduction pathways that turn on expression of genes with important functions in innate immunity. Among the proteins encoded by these genes are antimicrobial molecules such as antimicrobial peptides and interferons, chemokines and cytokines that recruit and activate other cells, enzymes such as iNOS (mentioned earlier) that generate antimicrobial molecules, and proinflammatory mediators (i.e., components that promote inflammation). Innate Immunity | CHAPTER 5 153 Cellular Pattern Recognition Receptors Activate Responses to Microbes and Cell Damage Several families of cellular PRRs contribute to the activation of innate immune responses that combat infections. Some of these PRRs are expressed on the plasma membrane, while others are actually found inside our cells, either in endosomes/lysosomes or in the cytosol. This array of PRRs ensures that the cell can recognize PAMPs on both extracellular and intracellular pathogens. DAMPs released by cell and tissue damage also can be recognized by both cell surface and intracellular PRRs. Many cell types in the body express these PRRs, including all types of myeloid white blood cells (monocytes, macrophages, neutrophils, eosinophils, mast cells, basophils, dendritic cells) and subsets of the three types of lymphocytes (B cells, T cells, and NK cells). PRRs are also expressed by some other cell types, especially those commonly exposed to infectious agents; examples include the skin, mucosal and glandular epithelial cells, vascular endothelial cells that line the blood vessels, and fibroblasts and stromal support cells in various tissues. While it is unlikely that any single cell expresses all of these PRRs, subsets of the receptors are expressed by various cell subpopulations. The rest of this section describes the four main families of mammalian PRRs and the signaling pathways that they activate, leading to protective responses. Toll-Like Receptors Recognize Many Types of Pathogen Molecules Toll-like receptors (TLRs) were the first family of PRRs to be discovered and are still the best-characterized in terms of their structure, how they bind PAMPs and activate cells, and the extensive and varied set of innate immune responses that they induce. Discovery of Invertebrate Toll and Vertebrate Toll-like Receptors In the 1980s, researchers in Germany discovered that Drosophila fruit fly embryos could not establish a proper dorsal-ventral (back to front) axis if the gene encoding the Toll membrane protein is mutated. (The name “Toll” comes from German slang meaning “weird,” referring to the mutant flies’ bizarrely scrambled anatomy.) For their subsequent characterization of the toll and related homeobox genes and their role in the regulation of embryonic development, Christiane NussleinVolhard, Eric Wieschaus, and Edward B. Lewis were awarded the Nobel Prize in Physiology and Medicine in 1995. But what does this have to do with immune system receptors? Many mutations of the toll gene were generated, and in 1996 Jules Hoffman and Bruno Lemaitre discovered that mutations in toll made flies highly susceptible to lethal infection with Aspergillus fumigatus, a fungus to which wild-type flies were immune (Figure 5-9). This striking observation led to other studies showing that Toll and related proteins are involved in the activation of innate immune responses in invertebrates. FIGURE 5-9 Impaired innate immunity in fruit flies with a mutation in the Toll pathway. Severe infection with the fungus Aspergillus fumigatus (yellow) results from a mutation in the signaling pathway downstream of the Toll pathway in Drosophila that normally activates production of the antimicrobial peptide drosomycin. [Electron micrograph adapted from B. Lemaitre et al., 1996, Cell 86:973; courtesy of J. A. Hoffman, University of Strasbourg.] For his pivotal contributions to the study of innate immunity in Drosophila, Jules Hoffman was a co-winner of the 2011 Nobel Prize in Physiology and Medicine. Characterization of the Toll protein surprisingly revealed that its cytoplasmic signaling domain was homologous to that of the vertebrate receptor for the cytokine IL-1 (IL-1R). Through a search for human proteins with cytoplasmic domains homologous to those of Toll and IL-1R, in 1997 Charles Janeway and Ruslan Medzhitov discovered a human gene for a protein similar to Toll that activated the expression of innate immunity genes in human cells. Appropriately, this and other vertebrate Toll relatives discovered soon thereafter were named Toll-like receptors (TLRs). Through studies with mutant mice, in 1998 Bruce Beutler obtained the important proof that TLRs contribute to normal immune functions in mammals. Mice homozygous for a mutant form of a gene called lps were resistant to the harmful responses induced by lipolysaccharide (LPS; also known as endotoxin), a major component of the cell walls of Gramnegative bacteria (Figure 5-10). In humans, a buildup of endotoxin from severe bacterial infection can induce too strong of an innate immune response, causing septic shock, a life-threatening condition in which vital organs such as the 154 PA R T I I (a) Gram negative bacteria E. coli | Innate Immunity Cell wall organization Lipopolysaccharide (endotoxin) Outer membrane Peptidoglycan Inner membrane (b) Gram positive bacteria S. aureus Cell wall organization Glycolipids Glycoproteins Peptidoglycan Membrane FIGURE 5-10 Cell wall components of Gram-negative and Gram-positive bacteria. Cell wall structures differ for Gramnegative (a) and Gram-positive (b) bacteria. Because of their thick peptidoglycan layer, Gram-positive bacteria retain the precipitate formed by the crystal violet and iodine reagents of the Gram stain, whereas the stain is easily washed out of the less-dense cell walls of Gram-negative bacteria. Thus the Gram stain identifies two distinct sets of bacterial genera that also differ significantly in other properties. [Photographs: E. coli photograph from Gary Gaugler/ brain, heart, kidney, and liver may fail. Each year, about 20,000 people die in the United States of septic shock caused by Gram-negative bacterial infections, so it was striking that some mutant strains of mice were resistant to fatal doses of LPS. Beutler found that the defective mouse lps gene encoded a mutant form of one TLR, TLR4, which differed from the normal form by a single amino acid so that it no longer was activated by LPS. This work provided an unequivocal demonstration that TLR4 is the cellular innate pattern recognition receptor that recognizes LPS and earned Beutler a share of the 2011 Nobel Prize. Thus, this landmark series of experiments showed in rapid succession that invertebrates respond to pathogens, that they use receptors also found in vertebrates, and that one of these receptors is responsible for LPS-induced innate immune responses. TLRs and Their Ligands Visuals Unlimited. S. aureus photograph from Dr. Fred Hossler/Getty Images] Intensive work over the last decade and a half has identified multiple TLR family members in mice and humans—as of 2011, 13 TLRs that function as PRRs have been identified in these species. TLRs 1-10 are conserved between mice and humans, although TLR10 is not functional in mice, while TLRs 11-13 are expressed in mice but not in humans. While TLRs have not been shown to be involved in vertebrate development, unlike in fruit flies, the set of TLRs present in a human or mouse can detect a wide variety of PAMPs from bacteria, viruses, fungi, and parasites, as well as DAMPs from damaged cells and tissues. Each TLR has a distinct repertoire of specificities for conserved PAMPs; the TLRs and some of their known PAMP ligands are listed in Table 5-4. Biochemical studies have revealed the structure of several TLRs and Innate Immunity TABLE 5-4 | CHAPTER 5 155 TLRs and their microbial ligands TLRs* Ligands Microbes TLR1 Triacyl lipopeptides Mycobacteria and Gram-negative bacteria TLR2 Peptidoglycans GPI-linked proteins Lipoproteins Zymosan Phosphatidlyserine Gram-positive bacteria Trypanosomes Mycobacteria and other bacteria Yeasts and other fungi Schistosomes TLR3 Double-stranded RNA (dsRNA) Viruses TLR4 LPS F-protein Mannans Gram-negative bacteria Respiratory syncytial virus (RSV) Fungi TLR5 Flagellin Bacteria TLR6 Diacyl lipopolypeptides Zymosan Mycobacteria and Gram-positive bacteria Yeasts and other fungi TLR7 Single-stranded RNA (ssRNA) Viruses TLR8 Single-stranded RNA (ssRNA) Viruses TLR9 CpG unmethylated dinucleotides Dinucleotides Herpes virus components Hemozoin Bacterial DNA Some herpesviruses Malaria parasite heme byproduct TLR10 Unknown Unknown TLR11 Unknown Profilin Uropathogenic bacteria Toxoplasma TLR12 Unknown Unknown TLR13 Unknown Vesicular stomatitis virus *All function as homodimers except TLR1, 2, and 6, which form TLR2/1 and TLR2/6 heterodimers. Ligands indicated for TLR2 bind to both; ligands indicated for TLR1 bind to TLR2/1 dimers, and ligands indicated for TLR6 bind to TLR2/6 dimers. how they bind their specific ligands. TLRs are membranespanning proteins that share a common structural element in their extracellular region called leucine-rich repeats (LRRs); multiple LRRs make up the horseshoe-shaped extracellular ligand-binding domain of the TLR polypeptide chain (Figure 5-11a). When TLRs bind their PAMP or DAMP ligands via their extracellular LRR domains, they are induced to dimerize. In most cases each TLR dimerizes with itself, forming a homodimer, but TLR2 forms heterodimers by pairing with either TLR1 or TLR6. How TLRs bind their ligands was not known until complexes of the extracellular LRR domain dimers with bound ligands were analyzed by x-ray crystallography. Structures of TLR2/1 with a bound lipopeptide and TLR3 with dsRNA are shown in Figure 5-11b; the characteristic “m”-shaped conformation of TLR dimers is apparent. As shown in Figure 5-12, TLRs exist both on the plasma membrane and in the membranes of endosomes and lysosomes; their cellular location is tailored to enable them to respond optimally to the particular microbial ligands they recognize. TLRs that recognize PAMPs on the outer sur- face of extracellular microbes (see Table 5-4) are found on the plasma membrane, where they can bind these PAMPs and induce responses. In contrast, TLRs that recognize internal microbial components that have to be exposed by the dismantling or degradation of endocytosed pathogens— nucleic acids in particular—are found in endosomes and lysosomes. Unique among the TLRs, TLR4 has been shown to move from the plasma membrane to endosomes/ lysosomes after binding LPS or other PAMPs. As we will see below, it activates different signaling pathways from the two locations. In addition to microbial ligands, TLRs also recognize DAMPs, endogenous (self) components released by dead/ dying cells or damaged tissues. Among the DAMPS recognized by plasma membrane TLRs are heat shock and chromatin proteins, fragments of extracellular matrix components (such as fibronectin and hyaluronin), and oxidized lowdensity lipoprotein and amyloid-. While self nucleic acids usually do not activate the intracellular PRR, under certain circumstances (such as when bound by anti-DNA or antichromatin antibodies in individuals with the autoimmune PA R T I I 156 | Innate Immunity (a) (b) Lipopeptide TLR2 TLR1 Leucine-rich repeats (LRRs) Exterior domain N N C Ribbon model of exterior domain C TLR3 TLR3' Cell membrane TIR domain N' N dsRNA C C' FIGURE 5-11 Toll-like receptor (TLR) structure and binding of PAMP ligands. (a) Structure of a TLR polypeptide chain. Each TLR polypeptide chain is made up of a ligand-binding exterior domain that contains many leucine-rich repeats (LRRs, repeating segments of 24-29 amino acids containing the sequence LxxLxLxx, where L is leucine and x is any amino acid), a membrane-spanning domain (blue), and an interior Toll/IL-1R (TIR) domain, which interacts with the TIR domains of other members of the TLR signal-transduction pathway. Two such polypeptide chains pair to form Toll/IL-1R (TLR) dimers, the form that binds ligands. (b) X-ray crystallographic structures of paired extracellular LRR domains of TLR dimers with bound ligands. Top: TLR2/1 dimer with a bound lipopeptide molecule. Bottom: TLR3/3 dimer with bound double-stranded (ds) RNA molecule. [Part (b) from Jin, M.S., and Lee, J-O. 2008. Immunity 29:182. PDB IDs 2Z7X (top) and 3CIY (bottom).] disease lupus erythematosus) they can be endocytosed and activate endosomal/lysosomal TLRs, contributing to the disease (see Chapter 16). that induce the expression of subsets of proteins, some of which are particularly effective in combating the type of pathogen recognized by the particular TLR(s). An important example is expression of the potent antiviral Type 1 interferons, IFNs- and -, induced by pathways downstream of the TLRs that bind viral components. As described below, activation of the interferon regulatory factors (IRFs) is essential for inducing transcription of the genes encoding IFN- and -. Combinations of transcription factors contribute to inducing the expression of many of these genes; examples include combinations of NF-B, IRFs, and/or transcription factors downstream of MAP kinase (MAPK) pathways (see Chapters 3 and 4), such as AP-1, that can be activated by signaling intermediates downstream of certain TLRs. The particular signal transduction pathway(s) activated by a TLR dimer following PAMP binding are largely determined by the TLR and by the initial protein adaptor (see Chapter 3) that binds to the TLR’s cytoplasmic domain. As shown in Figure 5-11a, this region is called the TIR domain (from Toll/IL-1 receptor), referring to the similarity noted earlier between the cytoplasmic domains of TLRs and IL-1 receptors (see Figure 4-5a). TIR domains of all TLR dimers serve as binding sites for the TIR domains of adaptors that activate the downstream Signaling Through TLRs: Overview Given the wide variety of potential pathogens the innate immune system needs to recognize and combat, how does binding of a specific pathogen evoke an appropriate response for that pathogen? Signaling through TLRs utilizes many of the principles and some of the signaling molecules described in Chapters 3 and 4, along with some unique to pathways activated by TLRs (and by other PRRs, described below). Detailed studies of the signaling pathways downstream of all of the TLRs, summarized schematically in Figure 5-13, have revealed that they include some shared components and activate expression of many of the same genes. An important example of a shared component is the transcription factor NF-B. NF-B is key for inducing many innate and inflammatory genes, including those encoding defensins; enzymes such as iNOS; chemokines; and cytokines such as the proinflammatory cytokines TNF-, IL-1, and IL-6, produced by macrophages and dendritic cells. There are also TLR-specific signaling pathways and components Innate Immunity Gram-negative bacteria Bacteria, parasites | CHAPTER 5 157 Gram-positive Flagellated Uropathogenic bacteria and bacteria bacteria fungi Plasma membrane TLR4/4 TLR2/1 TLR2/6 TLR5/5 TLR11/11 TLR4/4 TLR3/3 TLR7/7 TLR8/8 TLR9/9 Viral Viral proteins dsRNA Endosomes/lysosomes Viral ssRNA Viral ssRNA Bacterial/viral DNA FIGURE 5-12 Cellular location of TLRs. TLRs that interact with extracellular ligands reside in the plasma membrane; TLRs that bind ligands generated from endocytosed microbes are localized to endosomes and/or lysosomes. Upon ligand binding, the TLR4/4 dimer moves from the plasma membrane to the endosomal/lysosomal compartment, where it can activate different signaling components. signaling pathways. The two key adaptors are MyD88 (Myeloid differentiation factor 88) and TRIF (TIR-domain-containing adaptor-inducing IFN- factor). As shown in Figure 5-13, most TLRs, whether found on the cell surface or in endosomes/ lysosomes, bind the adaptor protein MyD88 (activating MyD88-dependent signaling pathways). In contrast, TLR3 binds the alternative adaptor protein TRIF (activating TRIFdependent signaling pathways). TLR4 is unique in binding both MyD88 (when it is in the plasma membrane, signaling its endocytosis) and TRIF (when it is in endosomes, after internalization). Figure 5-13 also shows the presence of two additional adaptors associated with most TLRs: TIRAP (TIR-domain containing adaptor protein) and TRAM (TRIF-related adaptor molecule). They are TIR domain-containing adaptors that serve as sorting receptors: TIRAP helps to recruit MyD88 to TLRs 2/1, 2/6, and 4, and TRAM helps to recruit TRIF to both TLR3 and endosomal/lysosomal TLR4. vated IKK then phosphorylates the inhibitory IB subunit of NF-B, releasing NF-B to enter the nucleus and activate gene expression. TAK1 does double duty in this TLR signaling cascade. After separating from the IKK complex, it activates MAPK signaling pathways (see Chapter 3) that result in the activation of transcription factors including Fos and Jun, which make up the AP-1 dimer (see Figure 3-16). In addition to activating NF-B and MAPK pathways via the MyD88-dependent pathway, the endosomal TLRs 7, 8, and 9, which bind microbial nucleic acids (especially viral RNA and bacterial DNA), also trigger pathways that activate IRFs. As shown in Figure 5-13, when triggered by these TLRs, the MyD88/IRAK4/TRAF6 complex activates a complex containing TRAF3, IRAK1, and IKK, leading to the phosphorylation, dimerization, activation, and nuclear localization of IRF7. IRF7 induces the transcription of genes for both Type 1 interferons, IFN- and -, which have potent antiviral activity. Thus different TLRs may differentially activate distinct transcription factors (NF-B, certain IRFs, and/or those activated by MAPK pathways), leading to variation in which genes are turned on to help protect us against the invading pathogens. MyD88-Dependent Signaling Pathways After associating with a TLR dimer following ligand binding, MyD88 initiates a signaling pathway that activates the NF-B and MAPK pathways by essentially the same pathway as that activated by IL-1 (see Figure 4-6). As shown for plasma membrane TLRs 2/1, 4, and 5 in Figure 5-13, MyD88 recruits and activates several IRAK protein kinases, which then bind and activate TRAF6. TRAF6 ubiquitinates NEMO and TAB proteins, leading to the activation of TAK1, which then phosphorylates the IB kinase (IKK) complex. Acti- TRIF-Dependent Signaling Pathways For the two endosomal TLRs that recruit the TRIF adaptor instead of MyD88—TLR3 and endosomal TLR4—the downstream signaling pathways differ somewhat from those activated by MyD88 (see Figure 5-13). TRIF recruits the RIP1 protein kinase that in turn recruits and activates TRAF6, 158 PA R T I I | Innate Immunity LPS TLR4/4 88 MyD P TIRA dsRNA TLR3/3 88 MyD P TIRA Endosome Lipopeptides TLR2/1 Flagellin TLR5/5 MyD88 Endosome Viral proteins TLR4/4 ssRNA TLR7/7 CpG DNA TLR9/9 IRAKs TRAF6 Ub TABs TAK1 RIP1 8 D8 F TRI M TRA 8 D8 Ub My My F TRI M TRA PI3K IRAK4 TRAF6 Ub TRAF6 Ub TRAF3 TBK-1 IKKε NEMO Ub P IKKα IKKβ P TRAF3 IRAK1 IKKα P P IRF7 IRF3 MAP kinase pathway NF-κB P IκB P IRF7 Ub Cytoplasm Nucleus P P IRF7 IRF3 NF-κB AP-1 P IRF7 IFN-β IFN-α Antimicrobials, cytokines, and chemokines FIGURE 5-13 TLR signaling pathways. Signaling pathways downstream of TLRs are shown for three cell surface TLRs (TLR2/1, TLR4, and TLR5; also activated by TLR2/6 and TLR 11 [not shown]) and for endosomal/lysosomal TLRs TLR7 and TLR 9 (in endosome on right; also activated by TLR8 [not shown]) and TLR3 and TLR4 (in endosome on left). MyD88-dependent signaling pathways are activated by the TIRAP and MyD88 receptor-binding adaptors. TRIF-dependent signaling pathways are activated by the TRAM and TRIF receptor-binding adaptors. Note: Not all signaling components are shown. See text for details. The MAPK pathway is shown in Figure 3-16. which then initiates the same steps as in the MyD88-dependent pathway. TLR3 also activates PI3K, which enhances MAPK pathway activation. In addition, TRIF activates TRAF3 and a complex of TBK-1 and IKK, which phosphorylates and activates IRF7 and a different IRF—IRF3. IRF3 and IRF7 dimerize and move into the nucleus and (together with NF-B and AP-1) induce the transcription of the IFN- and - genes. Thus, all of the intracellular TLRs that bind viral PAMPs in an infected cell induce the synthesis and secretion of Type I interferons, which feed back to potently inhibit the replication of the virus in that infected cell. C-Type Lectin Receptors Bind Carbohydrates on the Surfaces of Extracellular Pathogens The second family of cell surface PRRs that activate innate and inflammatory responses is the C-type lectin receptor (CLR) family. CLRs are plasma membrane receptors expressed Innate Immunity -1 tin CLR Viral dsRNA RLR 159 Fungal glucans c De Plasma membrane 2 CHAPTER 5 ure 5-14). Dectin-1 contains a cytoplasmic domain with an ITAM similar to those in the signaling chains of B-cell and T-cell antigen receptors (see Chapter 3). After Dectin-1 binds a ligand, its ITAM is phosphorylated and then recognized by the tyrosine kinase Syk, also involved in the initial stages of B-cell activation (see Figure 3-28). Syk triggers MAPK pathways, leading to the activation of the transcription factor AP-1, and also activates CARD9, one of many signaling components with Caspase recruitment domains (CARD; see Table 3-1 and additional examples below). CARD9 forms a complex with additional signaling components, leading to the activation of IKK and the nuclear translocation of NF-B as described above for TLR. NF-B and AP-1 cooperate in inducing expression of inflammatory cytokines and IFN-. Signals through different PRRs can modulate each other’s effects, enhancing or inhibiting expression of particular genes. As the signaling pathways downstream of some CLRs and TLRs are similar, signals coming into a cell from TLRs and variably on monocytes, macrophages, dendritic cells, neutrophils, B cells, and T-cell subsets. CLRs generally recognize carbohydrate components of fungi, mycobacteria, viruses, parasites, and some allergens (peanut and dust mite proteins). Humans have at least 15 CLRs that function as PRRs, most of which recognize one or more specific sugar moieties such as mannose (e.g., the mannose receptor and DC-SIGN), fucose (e.g., Dectin-2 and DC-SIGN), and glucans (e.g., Dectin-1). CLRs have a variety of functions. Unlike TLRs, which do not promote phagocytosis, some CLRs function as phagocytic receptors (see Table 5-3), and all CLRs trigger signaling pathways that activate transcription factors that induce effector gene expression. Some CLRs trigger signaling events that initially differ from TLR signaling but generally lead to downstream steps similar to the MyD88-dependent TLR pathways that activate the transcription factors NF-B and AP-1. An example of this is Dectin-1, which binds cell wall -1,3 glucans on mycobacteria, yeast, and other fungi (Fig- 1 | Syk P Peptidoglyan fragment 3 NLR Mitochondrion RIG-I NOD1 CARD9 MAVS RIP2 TRADD RIP1 Ub TRAF3 TBK1 IKKε P P IRF7 IRF3 NF-κB activation pathway MAP kinase pathway Cytoplasm Nucleus P P IRF7 IRF3 AP-1 NF-κB IFN-β IFN-α Antimicrobials, cytokines, and chemokines FIGURE 5-14 CLR, RLR, and NLR signaling pathways. Signaling pathways downstream of selected non-TLR PRRs are shown for (1) the plasma membrane CLR Dectin-1 (2) the cytosolic RLR RIG-I and (3) the cytosolic NLR NOD1. Note: Not all signaling components are shown. CARD domains are shown in brown. See text for details. The MAPK pathway is shown in Figure 3-16 and the NF-B pathway is described in Figure 5-13. 160 PA R T I I | Innate Immunity Dectin-1 (or related CLRs) that simultaneously recognize PAMPs on the same or different microbes can combine to enhance proinflammatory cytokine production. In addition, other signaling pathways downstream of Dectin-1 can activate and regulate different members of the NF-B family, which control expression of genes for additional cytokines that regulate helper-T-cell differentiation, as we will see later. Other CLRs have different cytoplasmic domains, downstream signaling pathways, and outcomes. For example, the binding of PAMPs to one CLR, DC-SIGN, on dendritic cells curtails inflammation by inducing expression of the anti-inflammatory cytokine IL-10 or by reducing mRNA levels for proinflammatory cytokines. In some cases such modulatory effects are beneficial, as responses can be switched from potentially harmful inflammatory responses to the activation of protective T-cell populations. However, as is discussed later, pathogens can also take advantage of these control circuits to reduce responses that would contribute to their elimination. Retinoic Acid-Inducible Gene-I-Like Receptors Bind Viral RNA in the Cytosol of Infected Cells The Retinoic acid-inducible gene-I-like receptors (RLRs) are soluble PRRs that reside in the cytosol of many cell types, where they play critical roles as sensors of viral infection (see Figure 5-14). The three known RLRs (RIG-I, MDA5, and LGP2) are CARD-containing RNA helicases that recognize viral RNAs, usually from RNA viruses such as influenza, measles, and West Nile. These receptors appear to distinguish viral from cellular RNA on the basis of particular structural features not shared by normal cellular RNA, such as doublestranded regions of the RNA, virus-specific sequence motifs, and, in the case of RIG-I, a 5 triphosphate modification that arises during viral RNA synthesis and processing. Upon viral RNA binding, RIG-I undergoes a conformational change that leads to binding via CARD-CARD interactions to its downstream adaptor molecule located in mitochondrial membranes, the mitochondria-associated MAVS (mitochondrial anti-viral signaling) protein. MAVS aggregates and recruits additional proteins, including the adaptor TRADD, TRAF3, and RIP1, which activate NEMO/IKKs that lead to the activation of IKK complexes, which activate IRFs 3 and 7 and the NF-B pathway, leading to expression of IFNs  and , other antimicrobials, chemokines, and proinflammatory cytokines. Nod-Like Receptors are Activated by a Variety of PAMPs, DAMPs, and Other Harmful Substances The final family of PRRs is the NLRs. (NLR is an acronym that stands for both Nod-like receptors and nucleotide oligomerization domain/leucine-rich repeat-containing receptors). The NLRs are a large family of cytosolic proteins activated by intracellular PAMPs and substances that alert cells to damage or danger (DAMPs and other harmful substances). They play major roles in activating beneficial innate immune and inflammatory responses, but, as we will see, some NLRs also trigger inflammation that causes extensive tissue damage and disease. The human genome contains approximately 23 NLR genes, and the mouse genome up to 34. NLR proteins are divided into three major groups, based largely on their domain structure, as shown in Figure 5-15: NLRCs (which have caspase recruitment domains, CARD), NLRBs (some of which have baculovirus inhibitory repeat, BIR, domains), and NLRPs (which have pyrin domains, PYD). The functions of many NLRs have not yet been well characterized; those of several NLRs are described below. NOD1 and NOD2 The best-characterized members of the NLRC family are NOD1 and NOD2. Important PAMPs recognized by NOD1 and NOD2 are breakdown products (such as muramyl dipeptides) produced during the synthesis or degradation of cell wall peptidoglycans of either intracellular or extracellular bacteria— peptides from the latter must enter the cell to activate NODs. Studies in mice have shown that NOD1 also provides protection against the intracellular protozoan parasite Trypanosoma cruzi, which causes Chagas disease in humans, and that NOD2 activates responses to some viruses, including influenza. PAMP binding to the LRR regions of NODs initiates signaling by activating NOD binding to the serine/threonine kinase RIP2 through their CARDs (Figure 5-14). RIP2 then activates IKK, leading to nuclear translocation of NF-B. RIP2 also activates MAPK pathways, leading to AP-1 activation. The activated NF-B and AP-1 initiate transcription of inflammatory cytokines—including TNF-, IL-1, and IL-6—and antimicrobial and other mediators. In addition, RIP2 activates the TRAF3/TBK1/IKK complex, leading to phosphorylation of IRFs 3 and 7 and to production of IFNs- and -. NLR Inflammasomes Whereas the NOD1 and NOD2 NLRCs activate the transcription of genes encoding inflammatory cytokines and antimicrobial and other proteins, some other NLRs do not trigger signaling pathways inducing expression of genes involved in innate immune responses. Instead, these NLRs assemble with other proteins into a complex that activates proteases necessary for converting the inactive large precursor forms (procytokines) of IL-1 and IL-18 into the mature forms that are secreted by activated cells. Because of the very potent inflammatory effects of secreted IL-1 (and also to some extent IL-18) (see Chapter 4), these complexes of certain NLRs with other proteins, including key proteases, are now referred to as inflammasomes. The discovery and surprising properties of inflammasomes are described in Advances Box 5-1. Expression of Innate Immunity Proteins is Induced by PRR Signaling The PRR-activated signaling pathways described above induce the transcription of genes encoding an arsenal of Innate Immunity | CHAPTER 5 161 NLRP1 CARD NOD1 NOD3 NOD4 NLRC4/ IPAF NOD2 NLRX1/ NOD5 FIIND NLRB NLRP3 NLRP4-14 LRR NAD NBD CARD BIR PYD NLRCs FIGURE 5-15 NLR families. NLRs are characterized by distinct protein domains. All NLRs have a leucine-rich repeat (LRR) domain, similar to those in TLRs, that functions in ligand binding in at least some NLRs, and a nuclear-binding domain (NBD). The three main classes are distinguished by their N-terminal domain (the lowest domains in the figure): NLRC receptors have CARD domains, NLRP proteins that help us to mount protective responses. Some of the induced proteins are antimicrobial and directly combat pathogens, while others serve key roles in activating and enhancing innate and adaptive immune responses. There is tremendous variation in what proteins are made in response to pathogens, reflecting their PAMPs as well as the responding cell types and their arrays of PRRs. Some of the most common proteins and peptides that are secreted by cells following PAMP activation of PRRs and that contribute to innate and inflammatory responses are listed in Table 5-5. Antimicrobial Peptides Defensins and cathelicidins were mentioned earlier as being important in barrier protection, such as on the skin and the epithelial layers connected to the body’s openings (see Table 5-2). Some cells and tissues constitutively (i.e., continually, without activation) express these peptides. For example, human intestinal Paneth epithelial cells constitutively express -defensins and some -defensins, and some defensins and the cathelicin LL-37 are constitutively synthesized and packaged in the granules of neutrophils, ready to NRLPs NLRB receptors have pyrin domains (PYD), and NLRB receptors have BIR (baculovirus inhibitor repeat) domains (although the related protein IPAF does not). These domains function in protein-protein interactions, largely through homotypic domain interactions. [Adapted from F. Martinon et al. 2007. Annual Review of Immunology 27:229; and E. Elinav et al. 2011. Immunity 34:665.] kill phagocytosed bacteria, fungi, viruses, and protozoan parasites. However, in some other cell types, such as mucosal and glandular epithelial cells, skin keratinocytes, and NK cells, the expression of these antimicrobial peptides is induced or enhanced by signaling through PRRs, in particular TLRs and the NOD NLRs. Macrophages do not produce these antimicrobial peptides following PRR activation; interestingly, however, there is an indirect pathway by which microbes induce cathelicidin in macrophages. Binding of microbial ligands to macrophage TLRs induces increased expression of receptors for vitamin D; binding of vitamin D to these receptors activates the macrophages to produce cathelicidin, which then can help the macrophages kill the pathogens. Type I Interferons Another major class of antimicrobial proteins transcriptionally induced directly by PRRs is the Type I interferons (IFN-,). Type I interferon production is generally activated by those cell surface TLRs and intracellular TLRs, RLRs, and NLRs that recognize viral nucleic acids and other 162 PA R T I I | Innate Immunity ADVANCES Inflammasomes The cytokine IL-1 has been recognized as one of the potent inducers of inflammation. While it was known that the IL-1 gene is transcriptionally activated following exposure to pathogen, damage, or danger-associated molecules, additional steps were needed to generate the mature IL-1 protein from its large intracellular pro-IL-1 precursor. An enzyme, initially called IL-1 Converting Enzyme (ICE), now known as caspase-1, was shown to carry out the cleavage of pro-IL-1, but the enzyme itself existed in most cells as a large inactive precursor. The breakthrough in understanding how mature IL-1 is produced came in 2002 when Jurg Tschopp and others published biochemical studies show- ing that activation of cells by LPS induced the formation of a large multiprotein aggregate containing an NLR and mature caspase-1 that cleaved pro-IL-1 into mature IL-1, allowing its release from cells. Because of the importance of IL-1 in promoting inflammation, Tschopp and his colleagues coined the term inflammasome for the large protein complex that activates caspase-1 to generate IL-1. Three NLRs (NLRP1, NLRP3, and NLRC4) have been shown to form inflammasomes that activate caspase-1 to cleave the large precursors of both IL-1 and IL-18, generating the mature proinflammatory cytokines. The best-characterized inflammasome is the NLRP3 inflammasome, which is expressed by monocytes, macrophages, neutrophils, dendritic cells, and some lymphocytes and epithelial cells. NLRP3 is a large complex containing multiple copies each of NLRP3, the adaptor protein ASC (which binds to NLRP3 via homotypic PYD:PYD domain interactions), and caspase-1 (Figure 1). NLRP3 can be activated in cells by a variety of components from bacteria, fungi, and some viruses. In addition to microbial components, NLRP3 can also be activated by nonmicrobial (“sterile”) substances, including several DAMPs released by damaged tissues and cells, such as hyaluronan, -amyloid (associated with Alzheimer’s plaques), and extracellular ATP and glucose. Recent NLRP3 NBD PYD NAD LRR PYD CARD Caspase-1 CARD ASC Pathogen activators Sterile activators Bacteria-derived Pore-forming toxins (S. aureus, Clostridia) Flagellin Peptidoglycan fragment RNA DNA Self-derived ATP Cholesterol crystals Urate crystals Glucose Amyloid β Hyaluronan Virus-derived RNA Influenza M2 protein Environment-derived Alum Asbestos Silica Alloy particles UV radiation Skin irritants Fungus-derived β-glucans Mannan Zymosan Hyphae Protozoa-derived Hemozoin IL-1 IL-18 FIGURE 1 The NLRP3 inflammasome and its activators. Assembly of the NLRs, ASC, and caspase-1 due to homotypic domain interactions leads to the formation of a pentamer or heptamer structure: the inflammasome. Activators of the inflammasome are divided into two categories: sterile activators include non-microbial host- and environment-derived molecules, and pathogen-associated activators include PAMPs derived Cell death from bacteria, virus, fungus, and protozoa. Activation of the inflammasome leads to maturation and secretion of IL-1 and IL-18 and, in some instances, to novel processes leading to inflammatory cell death. (Abbreviations: ASC, apoptosis-associated speck-like protein containing a caspase recruitment domain. For others, see legend to Figure 5-15 and text). [Adapted from B. K. Davis et al. 2011. Annual Review of Immunology 29:707.] Innate Immunity | CHAPTER 5 163 BOX 5-1 research has also implicated NLRP3 in mediating serious inflammatory conditions caused by an unusual class of harmful substances: crystals. Crystals of monosodium urate in individuals with hyperuricemia cause gout, an inflammatory joint condition, and inhalation of environmental silica or asbestos crystals cause the serious, often fatal, inflammatory lung conditions silicosis and asbestosis. When these crystals are phagocytosed, they damage lysosomal membranes, releasing lysosomal components into the cytosol. Similar effects may be responsible for the loosening (aseptic osteolysis) of artificial joints, caused by tiny metal alloy particles from the prosthesis that also activate NLRP3-mediated inflammation. How these disparate PAMPs, DAMPs, crystals, and metal particles all activate the NLRP3 inflammasome is under intense investigation. However, to date there is no conclusive evidence for the binding of any ligands directly to NLRP3; thus it may not act as a PRR for PAMPs and DAMPs per se but instead may sense changes in the intracellular milieu resulting from exposure to these materials. Current models, summarized in Figure 2, involve the activation by these varied stimuli of a common set of intracellular signals, such as potassium ion efflux, reactive oxygen species (ROS), and/or leakage of lysosomal contents, all of which appear to induce NLRP3 inflammasome assembly and caspase-1 activation, leading to the processing and secretion of IL-1 and IL-18. It is clear, however, that inflammasomes usually work in tandem with other PRRs in a cell to generate these cytokines. Activation of many TLRs, CLRs, RLRs, and NOD NLRs by PAMPs or DAMPs primes the inflammatory response by inducing transcription of the IL-1 and IL-18 genes by the signaling pathways described above. Once activated by as yet unknown intracellular mediators, the NLRP3 inflammasome then processes the procytokine precursors to the mature, active IL-1 and IL-18, which then are released by the cell by an unknown atypical secretion process. Somewhat less is known about the other inflammasomes, although they are also large multiprotein aggregates containing NLRs or related proteins and caspases and generate mature IL-1 and IL-18. The inflammasome discovered by Tschopp contains the NLR protein now called NLRP1 (see Figure 5-15), which is expressed in a wide array of hematopoietic and other cell types. Its activation by the lethal toxin of Bacillus anthracis contributes to the lethality of anthrax. The NLRC4 (also called IPAF) inflammasome, expressed predominantly in hematopoietic cells, is activated by certain Gram-negative bacteria and by the cytosolic presence of bacterial flagellin. A fourth inflammasome utilizes the AIM2 protein, which does not contain an NBD and hence is not an NLR, but does contain a PYD domain and hence can form inflammasomes with ASC and caspase-1. Interestingly, AIM2, which is induced by interferons, has been shown recently to bind cytosolic double-stranded DNA, such as from intracellular viruses and bacteria, and also to generate mature IL-1 and IL-18. The recent discovery of inflammasomes has provided the answer to one long-standing question: How is mature IL-1 generated in cells in response to innate and inflammatory stimuli? But other challenges remain, the most important being to identify the mechanism(s) by which inflammasomes are activated, so that their key caspase-1 enzyme becomes functional. In addition, as no functions are known yet for several other NLRs, more inflammasomes may yet be discovered. DAMP/ PAMP DAMP/PAMP K+ ATP P2X7 Pannexin-1 Crystalline/ particulate substances 1 2 3 Reactive oxygen species Lysosomal rupture NLRP3 inflammasome assembly Activated Caspase-1 Pro-IL-1 Pro-IL-18 IL-1 IL-18 “Priming” by proinflammatory stimuli e.g. TLR ligands, TNF, IL-1 IL-1 IL-18 FIGURE 2 Models for NLRP3 inflammasome activation. Three major models for NLRP3 inflammasome activation have been proposed, which may not be mutually exclusive. Model 1: The NLRP3 activator, ATP, triggers P2X7-dependent pore formation by the pannexin-1 hemichannel, allowing potassium ion efflux and entry into the cytosol of extracellular DAMPs and/or PAMPs which then directly engage NLRP3. Model 2: Crystalline or particulate NLRP3 agonists are engulfed, and their physical characteristics lead to lysosomal rupture. Released lysosomal contents, including enzymes, then activate NLRP3 inflammasome in the cytoplasm by unknown mechanisms; one possibility is cathepsin-B-dependent processing of a direct NLRP3 ligand. Model 3: All DAMPs and PAMPs, including ATP and particulate/crystalline activators, induce the generation of reactive oxygen species (ROS). A ROS-dependent pathway triggers NLRP3 inflammasome complex formation. Following these initial activation steps, caspase-1 clustering in the inflammasome activates caspase-1, which then cleaves pro-IL-1 and pro-IL-18 precursors cytokines, generating IL-1 and IL-18, which are then secreted. In some instances the death of the cell may also be induced. [Adapted from K. Schroder and J. Tschopp. 2010. Cell 140:821.] 164 PA R T I I TABLE 5-5 Antimicrobials Cytokines | Innate Immunity Secreted peptides and proteins induced by signaling through PRR Peptides/proteins Produced by Act on Immune/inflammatory effects Defensins and cathelicidin* Epithelia (e.g., oro/nasal, respiratory, intestinal, reproductive tracts; skin keratinocytes, kidney); NK cells Pathogens Inhibit, kill Monocytes, immature dendritic cells, T cells Mast cells Chemoattractant; activate cytokine production Activate degranulation Interferons -, - Virus-infected cells, macrophages, dendritic cells, NK cells Virus-infected cells NK cells Macrophages, T cells Inhibit virus replication Activate Regulate activity IL-1 Monocytes, macrophages, dendritic cells, keratinocytes, epithelial cells, vascular endothelial cells Lymphocytes Bone marrow Vascular endothelium Enhances activity Promotes neutrophil production Activates; increases vascular permeability Induces acute-phase response Fever Monocytes, macrophages, dendritic cells, NK cells, epithelial cells, vascular endothelial cells Lymphocytes Bone marrow IL-6 TNF- Monocytes, macrophages, dendritic cells, mast cells, NK cells, epithelial cells Liver Hypothalamus Liver Regulates activity Promotes hematopoiesis → neutrophils Activates; increases vascular permeability Induces acute-phase response Hypothalamus Fever Macrophages Vascular endothelium Liver Activates Activates, increases vascular permeability, fluid loss, local blood clotting Induces acute-phase response Hypothalamus Fever Vascular endothelium Tumors Cytotoxic for many tumor cells GM-CSF Macrophages, vascular endothelial cells Bone marrow Stimulates hematopoiesis → myeloid cells IL-12, IL-18 Monocytes, macrophages, dendritic cells Naïve CD4 T cells Induce TH1 phenotype, IFN- production Activate Naïve CD8 T cells, NK cells Chemokines IL-10 Macrophages, dendritic and mast cells; NK, T, B cells Macrophages, dendritic cells Antagonizes inflammatory response, including production of IL-12 and TH1 cells Example: IL-8 (CXCL8) chemokine** Macrophages, dendritic cells, vascular endothelial cells Neutrophils, basophils, immature dendritic cells, T cells Chemoattracts cells to infection site * Defensins and cathelicidin LL-37 vary among tissues in expression and whether constitutive or inducible. ** Other chemokines that are induced by PRR activation of cells in certain tissues, including various epithelial layers, may also specifically recruit certain lymphoid and myeloid cells to that site. See text, Chapter 14, and Appendix III. Innate Immunity | CHAPTER 5 165 Type I interferons exert their antiviral and other effects by binding to a specific receptor, frequently called IFNAR (IFNalpha receptor), expressed by most cell types (see Chapter 4). Similar to signaling pathways downstream of many cytokine receptors (see Table 4-3 and Figure 4-11), binding of IFN- or - activates IFNAR to recruit and activate specific JAKs (JAK1 and Tyk2), which then activate specific STATs (STATs 1 and 2; see Figure 5-16). STAT1/1 and STAT1/2 dimers then induce expression of various genes, including those for the three proteins that block viral replication: protein kinase R (PKR), which inhibits viral (and cellular) protein synthesis by inhibiting the elongation factor eIF2; 2,5-oligoadenlyate A synthetase, which activates the ribonuclease RNase L that degrades viral mRNA; and Mx proteins, which inhibit both the transcription of viral genes into mRNAs and the assembly of virus particles. Reflecting the potent antiviral activities of Type I interferons, they are used to treat some viral infections, such as hepatitis B and C. In addition to their key components and activate the IRF transcription factors. As summarized in Table 5-5, Type I interferons are produced in two situations. When virus infected, many cell types are induced to make IFN- and/or IFN- following binding of cytosolic viral PAMPs, such as nucleic acids, to their RLRs or NLRs (see Figure 5-14). In addition, many cells express cell surface TLRs recognizing viral PAMPs and/or internalize virus without necessarily being infected, allowing endosomal TLRs to recognize viral components. Signaling from these TLRs activates the IRFs and IFN-, production. One particular type of dendritic cell, called the plasmacytoid dendritic cell (pDC) because of its shape, is a particularly effective producer of Type I IFNs. The pDCs endocytose virus that has bound to various cell surface proteins (including CLRs such as DC-SIGN, which, for example, binds HIV). TLRs 7 and 9 in the endosomes then are activated by viral PAMPs (ssRNA and CpG DNA, respectively), leading to IRF activation. INF-α/β Cytosol P IFNAR1 IFNAR2 P TYK2 P JAK1 STAT2 P STAT1 P P Nucleus Dimerized STAT Transcription Protein kinase R 2’, 5’-oligo A synthetase Mx proteins GTP GDP + P dsRNA dsRNA Polymerized Mx proteins Oligo A P elF2α Inhibition of translation RNase L mRNA degradation FIGURE 5-16 Major antiviral activities induced by Type I interferons. Interferons- and - bind to the IFNAR receptor, which then recruits and activates the JAK1 and TYK2 protein kinases. They bind and phosphorylate STATs 1 and 2, which dimerize, enter the nucleus, and stimulate expression of three proteins that activate antiviral effects. Protein kinase R (PKR) binds viral dsRNA and Inhibition of virus transcription and assembly inhibits the activity of the eIF2 translation initiation factor. 2,5-oligoadenylate synthase synthesizes 2,5-oligoadenylate, which activates a ribonuclease, RNaseL, that degrades mRNAs. Mx proteins self-assemble into ringlike structures that inhibit viral replication and formation of new virus particles. [Adapted from Fig. 3-46, A. DeFranco et al. 2007. Immunity (Primers in Biology). Sunderland, MA: Sinauer Associates.] 166 PA R T I I | Innate Immunity roles in controlling viral infections, Type I interferons have other immune-related activities in that they activate NK cells and regulate activities of macrophages and T cells. Treatment with IFN- has been shown to have beneficial effects in some forms of multiple sclerosis, a T-cell-mediated autoimmune disease with inflammatory involvement, probably by inhibiting production of certain proinflammatory T-cellderived cytokines (see Chapter 16). Cytokines Among the proteins transcriptionally induced by PRR activation are several key cytokines, which—while not directly antimicrobial—activate and regulate a wide variety of cells and tissues involved in innate, inflammatory, and adaptive responses. As introduced in Chapter 4, cytokines function as the protein hormones of the immune system, produced in response to stimuli and acting on a variety of cellular targets. Several key examples of cytokines induced by PRR activation during innate immune responses are listed in Table 5-5, along with their effects on target cells and tissues. Three of the most important cytokines are IL-1, IL-6, and TNF-, the major proinflammatory cytokines that act locally on blood vessels and other cells to increase vascular permeability and help recruit and activate cells at sites of infection; they also have systemic effects (see below). IL-1, IL-6, and GM-CSF also feed back on bone marrow hematopoiesis to enhance production of neutrophils and other myeloid cells that will contribute to pathogen clearance. Activation of monocytes, macrophages, and dendritic cells through some TLRs also induces production of IL-12 and IL-18, cytokines that play key roles in inducing naïve helper T cells to become TH1 cells, in particular by inducing production of IFN-. As will be discussed in Chapter 11, the hallmark cytokine of TH1 cells is IFN-, which stimulates cell-mediated immunity and is an important macrophage-activating cytokine. Hence IL-12 and IL-18 are also considered proinflammatory. IL-10 is another important cytokine specifically induced by some TLRs in macrophages, dendritic cells, other myeloid cells, and subsets of T, B, and NK cells. IL-10 is anti-inflammatory, in that it inhibits macrophage activation and the production of proinflammatory cytokines by other myeloid cells. IL-10 levels increase over time and contribute to controlling the extent of inflammation-caused tissue damage. Chemokines These small protein chemoattractants (agents that induce cells to move toward higher concentrations of the agent) recruit cells into, within, and out of tissues (see Chapters 4 and 14 and Appendix III). Some chemokines are responsible for constitutive (homeostatic) migration of white blood cells throughout the body. Other chemokines, produced in response to PRR activation, have key roles in the early stages of immune and inflammatory responses in that they attract cells that contribute both to clearing the infection or damage and to amplifying the response. The first chemokine to be cloned, IL-8 (also called CXCL8), is produced in response to activation—by PAMPs, DAMPs, or some cytokines—of a variety of cells at sites of infection or tissue damage, including macrophages, dendritic cells, epithelial cells, and vascular endothelial cells. One of IL-8’s key roles occurs in the initial stages of infection or tissue damage; it serves as a chemoattractant for neutrophils, recruiting them to sites of infection. Other chemokines are specifically induced by PRR activation of epithelial cells in certain mucosal tissues and serve to recruit cells specifically to those sites, where they generate immune responses appropriate for clearing the invading pathogen. For example, B cells are recruited to the lamina propria, the lymphocyte-rich tissue under the intestinal epithelium, by two chemokines, CCL28 and CCL20, produced by intestinal epithelial cells activated by PAMP binding to their TLRs. These activated epithelial cells (as well as local dendritic cells activated by PAMPs) also produce cytokines that stimulate the B cells to produce IgA, the class of antibodies most effective in protecting against mucosal infections (see Chapter 13). Enzymes: iNOS and COX2 Among other genes activated in many cell types by PRRactivated signaling pathways are those for two enzymes that contribute importantly to the generation of antimicrobial and proinflammatory mediators: inducible nitric oxide synthase (iNOS) and cyclooxygenase 2 (COX2). As described earlier, the iNOS enzyme catalyzes an important step in the formation of nitric oxide, which kills phagocytosed microbes (see Figure 5-8). COX2, whose synthesis is induced by PRR activation in monocytes, macrophages, neutrophils, and mast cells, is key to converting the lipid intermediate arachidonic acid to prostaglandins, potent proinflammatory mediators. The next section provides an overview of the main processes of inflammatory responses initiated by innate immune responses. Inflammation will be covered in more detail in Chapter 15. Inflammatory Responses When the outer barriers of innate immunity—skin and other epithelial layers—are damaged, the resulting innate responses to infection or tissue injury can induce a complex cascade of events known as the inflammatory response. Inflammation may be acute (short-term effects contributing to combating infection, followed by healing)—for example, in response to local tissue damage—or it may be chronic (long term, not resolved), contributing to conditions such as arthritis, inflammatory bowel disease, cardiovascular disease, and Type 2 diabetes. The hallmarks of a localized inflammatory response were first described by the Roman physician Celsus in the first century AD as rubor et tumor cum calore et dolore (redness and swelling with heat and pain). An additional mark of inflammation added in the second century by the physician Galen is loss of function ( functio laesa). Today we know that | Innate Immunity these symptoms reflect an increase in vascular diameter (vasodilation), resulting in a rise of blood volume in the area. Higher blood volume heats the tissue and causes it to redden. Vascular permeability also increases, leading to leakage of fluid from the blood vessels, resulting in an accumulation of fluid (edema) that swells the tissue. Within a few hours, leukocytes also enter the tissue from the local blood vessels. These hallmark features of inflammatory responses reflect the activation of resident tissue cells—macrophages, mast cells, and dendritic cells—by PAMPs and DAMPs to release chemokines, cytokines, and other soluble mediators into the vicinity of the infection or wound. Recruited leukocytes are activated to phagocytose bacteria and debris and to amplify the response by producing additional mediators. Resolution of this acute inflammatory response includes the clearance of invading pathogens, dead cells, and damaged tissue; the activation of the systemic acute phase response and additional physiological responses, including the initiation of wound healing; and the induction of adaptive immune responses. However, if the infection or tissue damage is not resolved, it can lead to a chronic inflammatory state that can cause more local tissue damage and potentially have systemic consequences for the affected individual. CHAPTER 5 167 Inflammation Results from Innate Responses Triggered by Infection, Tissue Damage, or Harmful Substances When there is local infection, tissue damage, or exposure to some harmful substances (such as asbestos or silica crystals in the lungs), sentinel cells residing in the epithelial layer— macrophages, mast cells, and dendritic cells—are activated by PAMPs, DAMPs, crystals, and so on to start phagocytosing the offending invaders (Figure 5-17). The cells are also activated to release innate immunity mediators that trigger a series of processes that collectively constitute the inflammatory response. The recruitment of various leukocyte populations to the site of infection or damage is a critical early component of inflammatory responses. PRR signaling activates resident macrophages, dendritic cells, and mast cells to release the initial components of cellular innate immune responses, including the proinflammatory cytokines TNF-, IL-1, and IL-6; chemokines; prostaglandins (following the induced expression of the COX2 enzyme); and histamine and other mediators released by mast cells. These factors act on the vascular endothelial cells of local blood vessels, increasing 1 Tissue damage and bacteria cause resident sentinel cells to release Tissue chemoattractants and vasoactive damage factors that trigger a local increase in blood flow and capillary permeability. 4 Phagocytes and antibacterial substances destroy bacteria. Bacteria 2 3 Permeable capillaries allow an influx of fluid (exudate) and cells. Exudate (complement, anti-microbial proteins)) Neutrophils and other phagocytes migrate to site of inflammation (chemotaxis). Neutrophils and other phagocytes Extravasation Capillary FIGURE 5-17 Initiation of a local inflammatory response. Bacterial entry through wounds activates initial innate immune mechanisms, including phagocytosis by and activation of resident cells, such as macrophages and dendritic cells. Recognition of bacteria by soluble and cellular pattern recognition molecules initiates an inflammatory response that recruits antimicrobial substances and phagocytes (first neutrophils and then monocytes) to the site of infection. 168 PA R T I I | Innate Immunity vascular permeability and the expression of cell adhesion molecules (CAMs) and chemokines such as IL-8. The affected epithelium is said to be inflamed or activated. Cells flowing through the local capillaries are induced by chemoattractants and adhesion molecule interactions to adhere to vascular endothelial cells in the inflamed region and pass through the walls of capillaries and into the tissue spaces, a process called extravasation that will be described in detail in Chapter 14. Neutrophils are the first to be recruited to a site of infection where they enhance local innate responses, followed by monocytes that differentiate into macrophages that participate in pathogen clearance and help initiate wound healing. In addition to these key events at the site of infection or damage, the key cytokines made early in response to innate and inflammatory stimuli—TNF-, IL-1, and IL-6—also have systemic effects, which will be described in more detail in Chapter 15. They induce fever (a protective response, as elevated body temperature inhibits replication of some pathogens) by inducing COX-2 expression, which activates prostaglandin synthesis, as mentioned above. Prostaglandin E2 (PGE2) acts on the hypothalamus (the brain center controlling body temperature), causing fever. These three proinflammatory cytokines also act on the liver, inducing the acute phase response, which contributes to the resolution of the inflammatory response. Proteins of the Acute Phase Response Contribute to Innate Immunity and Inflammation During the 1920s and 1930s, before the introduction of antibiotics, much attention was given to controlling pneumococcal pneumonia. Researchers noted changes in the concentration of several serum proteins during the acute phase of the disease, the phase preceding recovery or death. The serum changes were collectively called the acute phase response (APR), and the proteins whose concentrations rise during the acute phase are still called acute phase response proteins (APR proteins). The physiological significance of some APR proteins is still not understood, but we now know that some contribute to the innate immune response to infection. The acute phase response (discussed more fully in Chapter 15) is induced by signals that travel through the blood from sites of injury or infection. The proinflammatory cytokines TNF-, IL-1, and IL-6 are the major signals responsible for induction of the acute phase response; they act on hepatocytes in the liver, inducing them to secrete APR proteins at higher levels. Among APR proteins are many components of the complement system, which contribute to both innate and adaptive immune responses, and other proteins that function as opsonins, enhancing phagocytosis (see Table 5-3). As was mentioned briefly earlier and will be described in more detail in Chapter 6, complement components and fragments contribute to pathogen inactivation and clearance in a variety of ways, including serving as opsonins. Several other proteins are produced at higher levels during the APR. Mannose-binding lectin (MBL), described earlier, is a collectin that recognizes mannose-containing molecular patterns on microbes and promotes phagocytosis by blood monocytes. When MBL binds to the surface of microorganisms it also initiates complement activation (see Chapter 6). Another important APR made in the liver is C-reactive protein (CRP), which belongs to a family of pentameric proteins called pentraxins that bind ligands in a calcium-dependent reaction. Among the ligands recognized by CRP are a polysaccharide found on the surface of pneumococcal bacteria and phosphorylcholine, which is present on the surface of many microbes. CRP is an opsonin and also activates a complement-mediated attack on the microbe. Circulating CRP levels are considered an indicator of the level of ongoing inflammation. Two other pentraxins, serum amyloid protein and PTX, similarly function as both opsonins and activators of the complement pathway. Also released into the blood at higher levels are the surfactant protein opsonins SP-A and SP-D, mentioned earlier as providing protection against lung infections, and a number of proteins that participate in or regulate the coagulation (clotting) pathway, such as fibrinogen. While most of these APR proteins are always present in the blood at low levels, their increased concentrations during the acute phase response provide enhanced protective functions during infections. With the combined defenses mounted by innate and inflammatory responses, together with those of the laterarising adaptive immune responses, most infections are eliminated. Immune and inflammatory responses generally are self-limiting, so once the pathogen and damaged tissue are cleared, inflammation usually diminishes and the tissues heal. However, persistence of pathogens (e.g., in tuberculosis) or other harmful substances (e.g., monosodium urate crystals in gout) can cause chronic inflammation and continuing tissue damage and illness. Natural Killer Cells In addition to the mechanisms of innate immunity described above, which largely are the responsibility of nonlymphoid cell types, a population of lymphocytes also recognizes components associated with pathogens, damage, or stress and generates rapid protective responses. Natural killer (NK) cells constitute a third branch of lymphoid cells, along with B and T lymphocytes of the adaptive immune system, all differentiating from the common lymphoid progenitor into three separate lineages (see Chapter 2). Unlike B and T lymphocytes, whose receptors have tremendous diversity for foreign antigens, NK cells express a limited set of invariant, nonrearranging receptors that enable the cells to be activated by indicators of infection, cancer, or damage that are expressed by other cells. Again unlike B and T cells, which require days of activation, proliferation, and differentiation to generate their protective antibody and cell-mediated Innate Immunity immune responses, NK cells are preprogrammed to respond immediately to appropriate stimuli, releasing from preformed secretory granules effector proteins that kill altered cells by inducing apoptosis. This mechanism of cell-mediated cytotoxicity is also carried out by cytotoxic T cells, which appear days later (see Chapter 13). NK-cell-mediated cytotoxic activity is enhanced by IFN- produced early during virus infections, an example of positive feedback regulation in innate immunity. In addition to releasing cytotoxic mediators, activated NK cells also secrete cytokines—the proinflammatory cytokines IL-6 and TNF-, as well as Type I IFNs and the Type II IFN, IFN-, a potent macrophage activator that also helps to activate and shape the adaptive response. Thus, along with the Type I IFNs produced by virusinfected cells or induced during innate responses by viral PAMPs, NK cells are an important part of the early innate response to viral infections (as well as to malignancy and other indicators of danger). These early innate responses control the infection for the days to week it takes for the adaptive response (antibodies and cytotoxic T cells) to be generated. How do NK cells sense that our cells have become infected, malignant, or potentially harmful in other ways? As will be described more fully in Chapter 13, NK cells express a variety of novel receptors (collectively called NK receptors). Members of one group serve as activating receptors (of which more than 20 have been described in humans and mice) that have specificity for various cell surface ligands that serve as indicators of infection, cancer, or stress. While one of these activating ligands in mice has been shown to be a protein component of a virus (mouse cytomegalovirus), most NK activating receptors apparently have specificity not for a pathogen-associated component but for proteins specifically up-regulated on infected, malignant, or stressed cells, which serve as danger signals perceived by NK cells. NK cells also express TLRs, and binding of PAMPs or DAMPs can add to the activation signals. As will be described later, to limit potential killing of normal cells in our body, NK cells also express inhibitory receptors that recognize membrane proteins (usually conventional MHC proteins) on normal healthy cells and inhibit NK-cellmediated cytotoxic killing of those cells. Many virus-infected or tumor cells lose expression of their MHC proteins and thus do not send these inhibitory signals. Receiving an excess of activating signals compared to inhibitory signals tells an NK cell that a target cell is abnormal, and the NK cell is activated to kill the target cell. Thus NK cells are part of our innate sensing mechanisms that provide immediate protection, in this case recognizing and eliminating our own body’s cells that have become harmful. Recent studies suggest that initial activation increases the population and/or activity of responding NK cells, which therefore would generate a greater response when exposed later to the same activating ligand(s). This resembles the immunological memory in the adaptive immune system and represents the best example of immunological memory in the vertebrate innate immune system. | CHAPTER 5 169 Regulation and Evasion of Innate and Inflammatory Responses Innate immune responses and the inflammatory responses they induce play key roles in clearing infections and healing infected or damaged tissues. The importance of some of the individual molecules involved in the generation of innate and inflammatory responses is dramatically demonstrated by the impact on human health of genetic defects and polymorphisms (genetic variants) that alter the expression or function of these molecules. Some of these defects and their clinical consequences are described in Clinical Focus box 5-2. As illustrated by these conditions, and by the many known roles (cited throughout this chapter) of innate and inflammatory mechanisms in protecting us against pathogens, these responses are essential to keeping us healthy. However, some other disorders show that innate and inflammatory responses can also be harmful, in that overproduction of various normally beneficial mediators and uncontrolled local or systemic responses can cause illness and even death. Therefore it is important that the occurrence and extent of innate and inflammatory responses be carefully regulated to optimize the beneficial responses and minimize the harmful responses. Innate and Inflammatory Responses Can Be Harmful To be optimally effective in keeping us healthy, innate and inflammatory responses should use their destructive mechanisms to eliminate pathogens and other harmful substances quickly and efficiently, without causing tissue damage or inhibiting the normal functioning of the body’s systems. However, this does not always occur—a variety of conditions result from excessive or chronic innate and inflammatory responses. The most dangerous of these conditions is sepsis, a systemic response to infection that includes fever, elevated heartbeat and breathing rate, low blood pressure, and compromised organ function due to circulatory defects. Several hundred thousand cases of sepsis occur annually in the United States, and mortality rates range from 20% to 50%, but sepsis can lead to septic shock, circulatory and respiratory collapse that has a 90% mortality rate. Sepsis results from septicemia, infections of the blood, in particular those involving Gram-negative bacteria such as Salmonella, although other pathogens can also cause sepsis. The major cause of sepsis from Gram-negative bacteria is the cell wall component LPS (also known as endotoxin), which as we learned earlier is a ligand of TLR4. As we have seen, LPS is a highly potent activator of innate immune mediators, including the proinflammatory cytokines TNF-, IL-1, and IL-6; chemokines; and antimicrobial components. Systemic infections activate blood cells including monocytes and neutrophils, vascular endothelial cells, and resident macrophages and other cells in the spleen, liver, and other CLINICAL FOCUS Genetic Defects in Components of Innate and Inflammatory Responses Associated with Disease In combination with our rapidly expanding understanding of the mechanisms by which innate and inflammatory responses contribute to disease susceptibility and resistance, in recent years advances in human genetics have helped to identify a number of genetic defects that confer greater susceptibility to infectious and inflammatory diseases. The adverse effects of mutations in genes encoding essential components of innate and inflammatory processes highlight the critical roles of these proteins in keeping us healthy. Since 2003, when the first mutations in innate immune components that predispose individuals to recurrent bacterial infections were discovered, a number of mutations interfering with the generation of protective innate immune responses have been identified. Two examples were mentioned earlier in this chapter: defects in NADPH oxidase, which cause chronic granulomatous disease, and MBL deficiencies, which predispose to respiratory infections. Also leading to defects in innate immunity are mutations in two proteins—MyD88 and IRAK4—required for the MyD88-dependent signaling pathway downstream of all TLRs except TLR3 (see Figure 5-13). Children with these defects suffer from severe invasive S. pneumococcus infections, some fatal, and are also susceptible to Staphylococcus aureus and Pseudomonas aeruginosa (Figure 1a). The MyD88 mutations completely prevent cytokine and chemokine induction by ligands for TLRs 2/1, 2/6, 5, 7, and 8. Not surprisingly, the effects of these mutations are less significant for TLR3 (which activates TRIF signaling pathways instead of MyD88) and TLR4 (which activates both MyD88 and TRIF signaling pathways). That the MyD88 mutations do not leave these children more susceptible to a wider variety of pathogens probably reflects the induction of protective immunity by other PRRs as well as by the adaptive immune system. In fact, the children become less susceptible to infections as they get older (see Figure 1b), consistent with the buildup of adaptive immunological memory to these pathogens. Other genetic defects with clinical consequences have been identified in the pathways by which Type 1 interferons (IFN,) are induced by viral nucleic acid PAMPs and then block virus replication in infected cells. As highlighted in Figure 2, mutations that completely or partially block these pathways (red symbols) have been found in TLR3 and other components of the pathway that induce IFN-,. Mutations have also been found in TYK2 and STAT1, key components that activate the (b) 100 100 Patients surviving (%) Patients without infection (%) (a) antiviral effects of the interferons in infected cells. Interestingly, these mutations were all discovered in children presenting with herpes simplex virus (HSV) encephalopathy, a severe HSV infection of the central nervous system. Cells in the CNS express TLR3, and it may be that the mutations in these pathways severely disable innate responses that are critical to protection against CNS infection by this virus. Children with TYK2 and STAT1 mutations are also very susceptible to other infections, especially with mycobacteria, probably because macrophages must be activated by IFN- (which also utilizes TYK2 and STAT1 in its signaling pathway) to be able to kill these intracellular bacteria. The final set of genetic defects associated with disease states to be discussed here involves the effects of genetic variants in NLRs (including inflammasomes) in promoting inflammatory diseases. Genome-wide genetic association studies have indicated that a number of allelic variants of TLRs and NLRs are associated with inflammatory disorders. Several variants are associated with inflammatory bowel disease, which includes ulcerative colitis and Crohn’s disease. The most impressive genetic association was of Crohn’s disease with mutations in NOD2 in or near its ligand-binding LRR region. As 75 50 25 0 0 10 20 30 Months 40 50 75 50 25 0 0 10 20 Years 30 40 FIGURE 1 Severe bacterial infection and mortality among 60 children with MyD88 or IRAK4 deficiencies. (a) Decline in the percentage of children with these deficiencies who are asymptomatic reveals the incidence of the first severe bacterial infection during the first 50 months of life. (b) Survival curve of children with deficiencies shows reduced mortality after 5 years of age. [Adapted from J-L. Casenova et al. 2011. Annual Review of Immunology 29:447; based on data from C. Picard et al. 2010. Medicine 89:403] 170 BOX 5-2 Dendritic cell Virus-infected cell Virus ER IFNAR Endosome TRIF TYK-2 TLR7,8,9 TLR3 Anti-viral effects MyD88 UNC-93B IRAK4 JAK1 TRAF3 TRAF3 NEMO TBK1 IKKε IKK STAT1 IRF3 IRF7 STAT2 NF-κB IRF9 IFN-α, β FIGURE 2 Genetic defects affecting the production and antiviral effects of IFN-␣,␤. This schematic shows some protein components involved in Type I IFN production by dendritic cells and the Type I IFN response in virus-infected cells. Proteins in which genetic mutations have been identified that result in defective functions and are associated with greater susceptibility to viral diseases are shown in red. Viruses are taken up by dendritic cells via specific receptors, and viral nucleic acids are detected by the various TLRs expressed in endosomes. Transport to endosomes of TLRs 3, 7, 8, and 9 is dependent on the ER protein UNC93B. Cytoplasmic signaling components activate NOD2 is activated by bacterial cell wall fragments, investigators have hypothesized that intestinal epithelial cells with the mutant NOD2 PRRs are unable to activate adequate protective responses to gut bacteria and/or to maintain appropriate balance between normal commensals and pathogenic bacteria, and that these defective innate immune functions contribute to Crohn’s disease pathogenesis. Consistent with this hypothesis, recent studies have found that intestinal Paneth cells from NOD2-defective individuals secrete reduced amounts of -defensins which, as mentioned earlier, are essential for maintaining normal commensal gut flora. Genetic variants of the NLRP3 inflammasome have also been shown to be transcription factors, including IRFs and NF-B, leading to synthesis and secretion of IFN-,. TLR3, UNC93B, TRAF3, and NEMO deficiencies are associated with impaired IFN production, particularly during herpesvirus infection. The binding of IFN-, to their receptor induces the phosphorylation of JAK1 and TYK-2, activating the signal transduction proteins STAT-1, STAT-2, and IRF9. This complex translocates as a heterotrimer to the nucleus, where it acts as a transcriptional activator, binding to specific DNA response elements in the promoter region of IFN-inducible genes. TYK-2 and STAT-1 deficiencies are associated with impaired IFN responses. [Adapted from J. Bustamante et al., 2008, Current Opinion in Immunology 20:39.] associated with Crohn’s disease and other inflammatory disorders. In fact, mutations in NLRP3 (originally called cryopyrin) have been show to be responsible for a set of autoinflammatory diseases (i.e., noninfectious inflammatory diseases affecting the body’s tissues) collectively known as CAPS (for cryopyrin-associated periodic fever syndromes); one example is NOMID (neonatal onset multisystem inflammatory disorder). These devastating syndromes include many signs of systemic inflammation, including fever, rashes, arthritis, pain, and inflammation affecting the nervous system, with adverse effects on vision and hearing. More than 70 inherited and novel mutations in NLRP3 associated with CAPS have been identified. Most are in the NRLP3 NBD domain, though some are in the LRR domain. What many have in common is their deregulating effect on NLRP3 activation of caspase-1, which may become constitutively active. Cells from NOMID patients have recently been shown to secrete higher levels of IL-1 and IL-18, both spontaneously and induced by PAMPs and DAMPs, promoting chronic inflammation. Another consequence of constitutive NLRP3 activation is the death of the activated cells, releasing DAMPs that lead to more inflammation. Fortunately, new therapeutic approaches that inhibit IL-1 activity seem to alleviate these symptoms in some patients. 171 172 PA R T I I | Innate Immunity tissues, to release these soluble mediators. They, in turn, systemically activate vascular endothelial cells, inducing them to produce cytokines, chemokines, adhesion molecules, and clotting factors that amplify the inflammatory response. Enzymes and reactive oxidative species released by activated neutrophils and other cells damage the vasculature; this damage, together with TNF--induced vasodilation and increased vascular permeability, results in fluid loss into the tissues that lowers blood pressure. TNF also stimulates release of clotting factors by vascular endothelial cells, resulting in blood clotting in capillaries. These effects on the blood vessels are particularly damaging to the kidneys and lungs, which are highly vascularized. High circulating TNF- and IL-1 levels also adversely affect the heart. Thus the systemic inflammatory response triggered by septicemia can lead to circulatory and respiratory failure, resulting in shock and death. As high levels of circulating TNF-, IL-1, and IL-6 are highly correlated with morbidity, considerable effort is being invested in developing treatments that block the adverse effects of these normally beneficial molecules. While not as immediately dangerous as septic shock, chronic inflammatory responses resulting from ongoing activation of innate immune responses can have adverse consequences for our health. For example, a toxin from Helicobacter pylori bacteria damages the stomach by disrupting the junctions between gastric epithelial cells and also induces chronic inflammation that has been implicated in peptic ulcers and stomach cancer. Also, increasing evidence suggests that the noninfectious DAMPs cholesterol (as insoluble aggregates or crystals) and amyloid  contribute, respectively, to atherosclerosis (hardening of the arteries) and Alzheimer’s disease. Other examples of harmful sterile (noninfectious) inflammatory responses discussed earlier— including gout, asbestosis, silicosis, and aseptic osteolysis— are induced, respectively, by crystals of monosodium urate, asbestos, and silica, and by metal alloy particles from artificial joint prostheses. These varied substances are all potent inflammatory stimuli due to their shared ability to activate the NLRP3 inflammasome, resulting in the release of the proinflammatory cytokines IL-1 and IL-18. Innate and Inflammatory Responses are Regulated Both Positively and Negatively As innate immune responses play essential roles in eliminating infections but also can be harmful when not adequately controlled, it is not surprising that many regulatory processes have evolved that either enhance or inhibit innate and inflammatory responses. These mechanisms control the induction, type, and duration of these responses, in most cases leading to the elimination of an infection without damaging tissues or causing illness. Positive Feedback Mechanisms Innate and inflammatory responses are increased by a variety of mechanisms to enhance their protective functions. To amplify the early stages of protective innate responses, two of the initial cytokines induced by PAMP or DAMP binding to PRRs—TNF- and IL-1—activate pathways similar to those downstream of TLRs (see Chapter 4) and hence induce more of themselves, an example of positive feedback regulation. In other words, a cell’s response to TNF- or IL-1 signaling can include production of more TNF- and IL-1. In a parallel fashion, in some cells Type I interferons can signal cells through the IFNAR receptor to produce more IFN. Other positive feedback mechanisms triggered by PRR activation include the activation of higher transcription rates of the genes for some TLRs, for the subunits of NF-B itself, and for NLRP3 and caspase-1. Collectively the increases in these proteins amplify IL-1 secretion, as production of IL-1 requires not only the transcription of its gene but also the processing of its large precursor protein by caspase-1, which is activated by NLRP3. Negative Feedback Mechanisms On the other side of the equation, as uncontrolled innate and inflammatory responses can have adverse consequences, many negative feedback mechanisms are activated to limit these responses. Several proteins whose expression or activity is increased following PRR signaling inhibit steps in the signaling pathways downstream of PRRs. Examples include production of a short form of the adaptor MyD88 that inhibits normal MyD88 function, the activation of protein phosphatases that remove key activating phosphate groups on signaling intermediates, and the increased synthesis of IB, the inhibitory subunit that keeps NF-B in the cytoplasm. The activation of these and other intracellular negative feedback mechanisms can lead cells to become less responsive, limiting the extent of the innate immune response. In a wellstudied example, when macrophages are exposed continually to the TLR4 ligand LPS, their initial production of antimicrobial and proinflammatory mediators is followed by the induction of inhibitors (including IB and the short form of MyD88) that block the macrophages from continuing to respond to LPS. This state of unresponsiveness, called LPS tolerance (or endotoxin tolerance), reduces the possibility that continued exposure to LPS from a bacterial infection will cause septic shock. Other feedback pathways inhibit the inflammatory effects of TNF- and IL-1. Each induces production of a soluble version of its receptor or a receptor-like protein that binds the circulating cytokine molecules (see Chapter 4), preventing them from acting on other cells. In addition, the antiinflammatory cytokine IL-10 is produced late in the macrophage response to PAMPs; it inhibits the production and effects of inflammatory cytokines and promotes wound healing. Another anti-inflammatory cytokine, TGF-, is also produced by macrophages, dendritic cells, and other cells, especially after PRR activation by apoptotic cells. The inhibitory effects of TGF- reduce the likelihood that DAMPs released by cells undergoing apoptosis will induce inflammatory responses. As a final example of a negative feedback Innate Immunity TABLE 5-6 | CHAPTER 5 173 Pathogen evasion of innate and inflammatory responses Type of evasion Examples Avoid detection by PRRs Proteobacteria flagellin has a mutation that prevents it from being recognized by TLR5. Helicobacter, Coxiella, and Legionella bacteria have altered LPS that is not recognized by TLR4. HTLV-1 virus p30 protein inhibits transcription and expression of TLR4. Several viruses (Ebola, influenza, Vaccinia) encode proteins that bind cytosolic viral dsRNA and prevent it from binding and activating RLR. Block PRR signaling pathways, preventing activation of responses Vaccinia virus protein A46R and several bacterial proteins have TIR domains that block MyD88 and TRIF from binding to TLRs. Several viruses block TBK1/IKK-activation of IRF3 and IRF7, required for IFN production. West Nile Virus NS1 protein inhibits NF-B and IRF transport into the nucleus. Yersinia bacteria produce Yop proteins that inhibit inflammasome activity; the YopP protein inhibits transcription of the IL-1 gene. Prevent killing or replication inhibition Salmonella and Listeria bacteria rupture the phagosome membrane and escape to the cytosol. Mycobacteria tuberculosis blocks phagosome fusion with lysosomes. Vaccinia virus encodes a protein that binds to Type I IFNs and prevents them from binding to the IFN receptor. Hepatitis C virus protein NS3-4A and Vaccinia virus protein E3L bind protein kinase R and block IFN-mediated inhibition of protein synthesis. loop, levels of circulating TNF- and IL-1 that could potentially induce harmful inflammatory responses act via the hypothalamus to induce the adrenal medulla to secrete glucocorticoid hormones, which have a variety of potent anti-inflammatory effects, including inhibiting the production of TNF- and IL-1 themselves. Pathogens Have Evolved Mechanisms to Evade Innate and Inflammatory Responses As it is advantageous to pathogens to evolve mechanisms that allow them to evade elimination by the immune system, many have acquired the ability to inhibit various innate and inflammatory signaling pathways and effector mechanisms that would clear them from the body. Most bacteria, viruses, and fungi replicate at high rates and, through mutation, may alter their components to avoid recognition or elimination by innate immune effector mechanisms. Other pathogens have evolved complex mechanisms that block normally effective innate clearance mechanisms. A strategy employed especially by viruses is to acquire genes from their hosts that have evolved and function as inhibitors of innate and inflammatory responses. Several strategies by which viruses evade immune responses triggered by interferons and other cytokines are listed in Table 4-5. Examples of mechanisms by which pathogens more generally avoid detection by PRRs, activation of innate and inflammatory responses, or elimination by those responses are described in Table 5-6. Interactions Between the Innate and Adaptive Immune Systems The many layers of innate immunity are important to our health, as illustrated by the illnesses seen in individuals lacking one or another component of the innate immune system (see the Clinical Focus box and other examples mentioned earlier in this chapter). However, innate immunity is not sufficient to protect us fully from infectious diseases, in part because many pathogens have features that allow them to evade innate immune responses, as discussed above. Hence the antigen-specific responses generated by our powerful adaptive immune system are usually needed to resolve infections successfully. While our B and T lymphocytes are the key producers of adaptive response effector mechanisms— antibodies and cell-mediated immunity—it is becoming increasingly clear that our innate immune system plays important roles in helping to initiate and regulate adaptive 174 PA R T I I | Innate Immunity immune responses so that they will be optimally effective. In addition, the adaptive immune system has co-opted several mechanisms by which the innate immune system eliminates pathogens, modifying them to enable antibodies to clear pathogens. The Innate Immune System Activates and Regulates Adaptive Immune Responses When pathogens invade our body, usually by penetrating our epithelial barriers, the innate immune system not only reacts quickly to begin to clear the invaders but also plays key roles in activating adaptive immune responses. While innate cells at the infection site (resident macrophages, dendritic cells, mast cells, and newly recruited neutrophils) are sensing the invading pathogens through their pattern recognition receptors and generating antimicrobial and proinflammatory responses that will slow down the infection, they are also initiating steps to bring the pathogens to the attention of lymphocytes and help activate responses that, days later, will generate the strong antigen-specific antibody and cell-mediated responses that will resolve the infection. The first step in the generation of adaptive immune responses to pathogens is the delivery of the pathogen to lymphoid tissues where T and B cells can recognize it and respond. As is discussed in more detail in Chapter 14, dendritic cells (usually immature) that serve as sentinels in epithelial tissues bind microbes through various pattern-recognition receptors. The dendritic cells carry the bound microbes— either still attached to the cell surface or in phagosomes—via the lymphatic vessels to nearby secondary lymphoid tissues, such as the draining lymph nodes. There the dendritic cells can transfer or present the microbes or microbial components to other cells. In many cases the dendritic cell has internalized and degraded the microbe, and microbe-derived peptides come to the cell surface bound to its MHC class II proteins. Meanwhile, the binding of microbial PAMPs to the dendritic cell’s PRRs also has activated the dendritic cell to mature. It now expresses higher levels of MHC class II proteins and has turned on expression of costimulatory membrane proteins, such as CD80 or CD86 (see Chapter 3), that are recognized by receptors on TH cells. As a result of these processes of microbe binding, processing, and maturation, mature dendritic cells are the most effective antigen-presenting cells, particularly for the activation of naïve (not previously activated) TH cells. Pathogen-Specific Activation of TH Cell Subsets Activation of dendritic cells by binding of various PAMPs to PRRs has additional important consequences for adaptive immune responses. Depending on the nature of the pathogen and what PRRs and downstream signaling pathways are activated, dendritic cells are induced to secrete specific cytokines that influence what cytokines a naïve TH (CD4 ) cell will produce after activation, thus determining its functional phenotype (Figure 5-18). These distinct phenotypes, or TH cell subsets, will be discussed in detail in Chapter 11. As one example, a Gram-negative bacteria may express PAMPs that bind to TLR4, TLR5, and/or the endosomal/lysosomal TLRs that bind bacterial nucleic acids. This binding stimulates dendritic cells to produce IL-12, which induces naïve TH cells to become TH1 cells (see Chapter 4). The hallmark cytokine of TH1 cells is IFN-, which activates macrophages to kill intracellular bacteria, reinforces the TH1 phenotype, contributes to the activation of virus-specific cytotoxic T cells, and, along with IL-2 (another TH1 cytokine), helps to activate B cells. In contrast, binding of PAMPs from fungi, Gram-positive bacteria, or helminths to TLR 2/1 or 2/6 activates production of IL-10 and inhibits secretion of IL-12. This combination of effects of stimulation through TLR2-containing receptors supports naïve TH cell differentiation into TH2 cells, which promote strong antibody responses that are more important for providing protection against these pathogens than are cell-mediated responses. Fungal PAMPs bind to the CLR Dectin-1, which activates the dendritic cell to produce the proinflammatory cytokines TNF-, IL-6, and IL-23; the latter two induce the TH to become TH17 cells, which are characterized by the secretion of proinflammatory cytokines such as IL-17 and IL-6. These cytokines help recruit neutrophils to the site, which then phagocytose and kill the fungi. It is important to recognize that the schematic shown in Figure 5-18 is oversimplified, as there are many complexities to the regulation of TH phenotype not shown. For example, IL-4, a key cytokine for inducing formation of TH2 cells, may be made by mast cells and/or basophils following PRR activation by certain pathogens, including helminth worms. The resulting TH2 cells induce production of IgG and IgE antibodies that are particularly effective against some of these pathogens. In summary, through their differential activation of specific PRRs, various pathogens induce dendritic and other innate cells to produce cytokines that activate TH cells to acquire distinct cytokine-producing phenotypes. These distinct TH subsets generally elicit the types of adaptive immune responses that will be most effective in eliminating the particular pathogen invaders. The molecular mechanisms by which naïve CD4 TH cells are influenced to differentiate into various mature TH subsets will be described in Chapter 11. Pathogen-Specific Activation of TC Cells Recent research has also revealed key roles of dendritic cells in the activation of cytotoxic T-cell-mediated responses that are needed to kill cells in our body that have become harmful through infection with viruses or other intracellular pathogens or through malignant transformation. As was the case with naïve TH (CD4 ) cells, naïve cytotoxic TC (CD8 ) cells also are most effectively activated by mature dendritic cells (see Chapter 13). Again, dendritic cell maturation (also referred to as licensing) is usually induced by the binding of microbial components to a dendritic cell’s PRRs. These mature/licensed DCs also are capable of novel pathways for processing internalized | Innate Immunity CHAPTER 5 175 Bacteria Naïve TLR4 or TLR5 Dectin-1 T H1 IFN-γ IL-12 TLR3, 7, 9 Fungi IL-6 IL-23 Virus Naïve TH17 IL-17 IL-10 TLR2/1 Naïve Helminth TH2 IL-4 IL-5 IL-13 TLR2/6 Fungi IL-10 RA TGF-β CD28 CD80/86 TCR MHC II with peptide FIGURE 5-18 Differential signaling through dendritic cell PRRs influences helper T cell functions. Microbial PAMPs bind and activate distinct PRRs and signaling pathways that differentially induce production of various cytokines and other mediators, such as retinoic acid (RA). These cytokines interact with receptors on naïve CD4 T cells that are in the process of being activated both by antigen-derived peptides bound to dendritic cell MHC class II proteins pathogens so that peptides generated by endosomal/lysosomal degradation are carried to the cell surface bound to MHC class I proteins rather than to class II proteins (this process is called cross-presentation and explained in Chapter 8). These peptide/ MHC class I complexes, together with costimulatory proteins such as CD80 or CD86 on the surface of the mature/licensed dendritic cell, activate the naïve TC to give rise to cytotoxic T effector cells. As was the case with the activation of TH cells, cytokines such as IL-12 and IFN- produced by innate cells (dendritic cells, macrophages, NK cells, or other cells) contribute to the differentiation and activation of TC cells. T-Independent Antigens In addition to their indirect roles in promoting and regulating antibody and cell-mediated responses via their activation of DC and other innate cells, described above, TLRs also can be more directly involved in activating B and T cells. B cells express TLRs, and the binding of PAMPs to these TLRs activates signaling pathways that can add to or substitute for the signals from TH cells normally required for B-cell activation. One example has been well studied in mouse B cells. In combination with signals from the B cells’ BCR after antigen binding, TLR4 binding of LPS (at low concentrations) can activate sufficient signals to induce the B cell to proliferate Naïve Treg IL-10 TGF-β and by interactions with costimulatory molecules such as CD80 or CD86. Note that activation of dendritic cells by PAMP binding to TLRs stimulates increased expression of MHC class II and costimulatory proteins. The particular cytokine(s) to which a naïve CD4 T cell is exposed induces that cell turn on genes for certain cytokines, determining that cell’s functional phenotype (see text). [Adapted from B. Pulendran et al. 2010. Nature Immunology 11:647.] and differentiate into antibody-secreting plasma cells without TH cells. At high concentrations of LPS, TLR4-activated signals are sufficient to activate all B cells (polyclonal activation), regardless of their antigen-binding specificity; hence LPS has for many years been called a T-independent antigen (see Chapter 12). Human B cells do not express TLR4 and hence do not respond to LPS; however, they do express TLR9 and can be activated by microbial CpG DNA. The ability of TLR signals to replace TH signals is often beneficial; co-binding of bacteria to both BCR and TLR on a cell may activate that cell more quickly than if it had to wait for signals from a TH cell. Some T cells also express TLRs, which similarly function as costimulatory receptors to enhance protective responses. For example, mouse CD8 T cells express TLR2; costimulation of CD8 T cells by TLR2 ligands along with TCR recognition of peptide/MHC class I complexes reduces the T cells’ need for costimulatory signals provided by dendritic cells and enhances their proliferation, survival, and functions. Adjuvants Activate Innate Immune Responses to Increase the Effectiveness of Immunizations Given these activating and potentiating effects of PRR ligands on adaptive immune responses, can they be used to enhance 176 PA R T I I | Innate Immunity the efficacy of vaccines in promoting protective immunity against various pathogens? In fact, many of the materials— known as adjuvants—that have been shown over the years by trial and error to enhance immune responses in both laboratory animals and humans contain ligands for TLRs or other PRRs (see Chapter 17). For example, Complete Freund’s Adjuvant, perhaps the most potent adjuvant for immunizations in experimental animals, is a combination of mineral oil and killed Mycobacteria. The mineral oil produces a slowly dispersing depot of antigen, a property of many effective adjuvants, while fragments of the bacteria’s cell wall peptidoglycans serve as activating PAMPs. Alum (a precipitate of aluminum hydroxide and aluminum phosphate) is used in some human vaccines; it has recently been shown to activate the NRLP3 inflammasome, thereby enhancing IL-1 and IL-18 secretion and promoting inflammatory processes that enhance adaptive immune responses. As the responding innate cells do not make IL-12, immunizations with alum usually lead the activated T cells to become TH2 cells, which promote strong antibody responses. While many vaccines consist of killed or inactivated viruses or bacteria and hence contain their own PAMPS, which function as built-in adjuvants, some new vaccines consist of protein antigens that themselves are not very stimulatory to the immune system. Tumor antigens also tend to induce weak responses. Hence, considerable effort is being invested in developing new adjuvants based on current knowledge about PRRs. LPS is a highly potent adjuvant but generates too much inflammation to use; less harmful versions of LPS are being developed as potential adjuvants. The ability of the TLR3 ligand poly I:C (synthetic double-stranded RNA) to activate innate immunity and initiate inflammatory responses has inspired clinical trials to test its ability to enhance the effectiveness of weak vaccines, including antitumor vaccines. Vaccines using the TLR9 ligand CpG DNA (which mimics bacterial DNA) as an adjuvant are in clinical trials; there is great interest in CpG DNA as it preferentially elicits a TH1 response, important for inducing cell-mediated immunity (see Figure 5-18). Other approaches for generating an effective vaccine for a pathogen protein include fusing the protein to a TLR ligand using genetic engineering. For example, fusions of pathogen proteins to the TLR5 ligand flagellin are currently being tested. These are all examples of how our expanding knowledge of the interactions between the innate and adaptive immune systems may contribute to the development of more effective vaccines. Some Pathogen Clearance Mechanisms are Common to Both Innate and Adaptive Immune Responses Adaptive immune responses—antibody responses in particular—have adopted and modified several effector functions by which the innate immune system eliminates antigen, so that they are also triggered by antibody binding to antigens. While some will be discussed in more detail in Chapters 6 and 13, several examples are mentioned briefly here as illus- trations of important interactions between innate and adaptive immune responses. As discussed earlier in this chapter, several soluble proteins that recognize microbial surface components—including SP-A, SP-D, and MBL—function as opsonins; when they are bound to microbial surfaces they are recognized by receptors on phagocytes, leading to enhanced phagocytosis. Antibodies of the IgG and IgA classes also serve as opsonins; after binding to microbial surfaces these antibodies can be recognized by Fc receptors that are expressed on macrophages and other leukocytes, triggering phagocytosis (see Chapter 3). One receptor for IgG Fc regions (FcR), called CD16, is also expressed on NK cells, where it serves as an activating receptor. Thus, when IgG antibodies bind to foreign antigens on the cell surfaces (such as viral or tumor antigens), the IgG antibodies can be recognized by the FcR on the NK cell, triggering NK-cell-mediated cytotoxic killing of the infected or malignant cell (see Chapter 13). Finally, as mentioned earlier and described fully in Chapter 6, the complement pathway can be activated by both innate and antibody-mediated mechanisms. Components on microbe surfaces can be directly recognized by soluble pattern-recognition proteins, including MBL and C-reactive protein, leading to activation of the complement cascade. Similarly, when antibodies of the IgM class and certain IgG subclasses bind microbe surfaces, their Fc regions are recognized by the C1 component of complement, also triggering the complement cascade. As was shown in Figure 5-7, MBL and C1 are related structurally. Once the complement pathway is activated by any of these microbe-binding proteins, it generates a common set of protective activities; various complement components and fragments promote opsonization, lysis of membrane-bounded microbes, and generation of fragments that have proinflammatory and chemoattractant activities. Thus, the adaptive immune system makes good use of mechanisms that initially evolved to contribute to innate immunity, co-opting them for the elimination of pathogens. Ubiquity of Innate Immunity Determined searches among plant and invertebrate animal phyla for the signature proteins of the highly efficient adaptive immune system—antibodies, T cell receptors, and MHC proteins—have failed to find any homologs of these important vertebrate proteins. Yet without them multicellular organisms have managed to survive for hundreds of millions of years. The interior spaces of organisms as diverse as the tomato, fruit fly, and sea squirt (an early chordate, without a backbone) do not contain unchecked microbial populations. Careful studies of these and many other representatives of nonvertebrate phyla have found arrays of well-developed processes that carry out innate immune responses. The accumulating evidence leads to the conclusion that multiple immune mechanisms protect all multicellular organisms Innate Immunity TABLE 5-7 | CHAPTER 5 177 Immunity in multicellular organisms Taxonomic group Higher plants Innate immunity (nonspecific) Adaptive immunity (specific) Invasioninduced protective enzymes and enzyme cascades  Phagocytosis  Anti microbial peptides Pattern recognition receptors Lymphocytes Variable lymphocyte receptors Antibodies    Invertebrate animals Porifera (sponges)  ?    Annelids (earthworms)  ?    Arthropods (insects, crustaceans)     Vertebrate animals Jawless fish (hagfish, lamprey)  Elasmobranchs (cartilaginous fish; e.g., sharks, rays)  Bony fish (e.g., salmon, tuna)  Amphibians  Reptiles  Birds  Mammals  Sources: M. J. Flajnik and L. Du Pasquier, 2008, “Evolution of the Immune System,” in Fundamental Immunology (6th ed.), W. E. Paul, ed., Philadelphia: Lippincott; J. H. Wong et al., 2007, “A review of defensins of diverse origins,” Current Protein and Peptide Science 8:446. from microbial infection and exploitation (Table 5-7). Some of the innate immune system components occur across the plant and animal kingdoms. For example, as mentioned early in this chapter, virtually all plant and animal species, and even some fungi, have antimicrobial peptides similar to defensins. Most multicellular organisms have patternrecognition receptors containing leucine-rich repeats (LRRs), although many organisms also have other families of PRRs. While the innate immune responses activated by these receptors in plants and invertebrates show both similarities and differences compared to those of vertebrates, innate immune response mechanisms are essential for the health and survival of these varied organisms. Plants Rely on Innate Immune Responses to Combat Infections Despite plants’ tough outer protective barrier layers, such as bark and cuticle, and the cell walls surrounding each cell, plants can be infected by a wide variety of bacteria, fungi, and viruses, all of which must be combated by the plant innate immune system. Plants do not have phagocytes or other circulating cells that can be recruited to sites of infection to mount protective responses. Instead, they rely on local innate immune responses for protection against infection. As described in Box 5-3, some resemble innate responses of animals, while others are quite distinct. Invertebrate and Vertebrate Innate Immune Responses Show Both Similarities and Differences Additional innate immune mechanisms evolved in animals, and we vertebrates share a number of innate immunity features with invertebrates. PRRs (including relatives of Drosophila Toll and vertebrate TLRs) that have specificities for microbial carbohydrate and peptidoglycan PAMPs are found in organisms as primitive as sponges. Together with soluble opsonin proteins (including some related to complement components), some of these early invertebrate PRRs function in promoting phagocytosis, an early mechanism for clearing pathogens. Innate signaling has been well studied in Drosophila, and signaling proteins have been identified in the flies that are homologous to several of those downstream of vertebrate TLRs (including MyD88 and IRAK homologs). A difference is that fly Toll does not bind to PAMPs directly but is indirectly activated by the soluble protein product of an enzyme cascade triggered by pathogen binding to soluble PA R T I I 178 | Innate Immunity EVOLUTION Plant Innate Immune Responses In the plasma membrane under the cell wall, plant cells express pattern recognition receptors with LRR domains reminiscent of animal TLRs. These PRRs recognize what plant biologists refer to as microbe-associated molecular patterns (MAMPs), including bacterial flagellin, a highly conserved bacterial translation elongation factor, and various bacterial and fungal cell wall components (Figure 1). As is true for animal TLRs, some plant PRRs respond to danger-associated molecular patterns (DAMPs), which usually are created by pathogen enzymes that attack and fragment cell wall components. Some bacterial and fungal pathogens directly inject into plant cells toxin effector proteins that inhibit signaling through the plasma membrane PRRs. These toxins are recognized by a distinct class of LRR receptors in the cytoplasm called R proteins, which, like animal NLR proteins, have both LRR and nucleotide-binding domain domains. After ligand binding, plant PRRs activate signaling pathways and transcription factors distinct from those of Oomycetes Fungi MAMPs DAMPs Bacteria PRRs Plant cell wall PRRs Cell wall degrading enzymes Extracellular effectors PRRs Intracellular effectors Cytoplasm Innate immune responses R-protein PRRs FIGURE 1 Activation of plant innate immune responses. Microbe-associated molecular patterns (MAMPs), danger-associated molecular patterns (DAMPs) generated by microbial enzymes (such as by degrading the cell wall), and microbial effectors (such as toxins) are recognized by plasma membrane LRR-containing pattern recognition receptors (PRRs). Microbial effectors that enter the cytoplasm are recognized by a class of PRRs called resistance (R) proteins. Recognition of MAMPs, DAMPs, and effectors by PRRs induces innate immune responses. [Adapted from T. Boller and G. Felix, 2009, Annual Review of Plant Biology 60:379.] | Innate Immunity CHAPTER 5 179 BOX 5-3 vertebrate cells (plants do not have NF-B or IRF homologs), triggering innate responses. The primary protective innate immune response mechanisms of plants to infection are the generation of reactive oxygen and nitrogen species, elevation of internal pH, and induction of a variety of antimicrobial peptides (including defensins) and antimicrobial enzymes that can digest the walls of invading fungi (chitin- ases) or bacteria (-1,3 glucanase). Plants may also be activated to produce organic molecules, such as phytoalexins, that have antibiotic activity. In some cases, the responses of plants to pathogens even goes beyond these protective substances to include structural responses. For example, to limit infection of leaves, PAMP binding to PRRs induces the epidermal guard cells that form the openings (stomata) involved in leaf gas exchange to Tomato close, preventing further invasion (Figure 2). Other protective mechanisms include the isolation of cells in the infected area by strengthening the walls of surrounding noninfected cells and the induced death (necrosis) of cells in the vicinity of the infection to prevent the infection from spreading to the rest of the plant. Mutations that disrupt any of these processes usually result in loss of the plant’s resistance to a variety of pathogens. Tobacco Open stomata PAMPs or bacteria PAMPs or bacteria Closed stomata FIGURE 2 Induced closure of leaf stomata following exposure to bacterial PAMPs. In normal light conditions, the openings (stomata) formed by pairs of guard cells in the leaf epidermis are open (visible as the football-shaped openings in the upper pair of leaf photographs). However, as shown in the lower panels, exposure of the tobacco leaf to the bacterial plant pathogen Pseudomonas syringae and exposure of the tomato leaf to bacterial LPS induce the closure of the stomata (no openings visible). [From M. Melotto et al. 2008. Annual Review of Phytopathology 46:101–122.] 180 PA R T I I | Innate Immunity pattern-recognition proteins. Through this pathway bacterial and fungal infections lead to the degradation of an IB homolog and activation of NF-B family members, Dif and Dorsal, which induce the production of drosomycin, an insect defensin, and other antimicrobial peptides. In addition to these and other pathways activated by PRRs, Drosophila and other arthropods employ other innate immune strategies, including the activation of phenoloxidase cascades that result in melanization—the deposition of a melanin clot around invading organisms that prevents their spread. Thus invertebrates and vertebrates have common as well as distinct innate immune response mechanisms. S U M M A R Y ■ ■ ■ ■ ■ ■ The receptors of the innate immune system recognize conserved pathogen-associated molecular patterns (PAMPs), which are molecular motifs found in microbes, and damage-associated molecular patterns (DAMPs) from aging, damaged, or dead cells. Therefore these receptors are called pattern-recognition receptors (PRRs). The epithelial layers that insulate the interior of the body from outside pathogens—the skin and epithelial layers of the mucosal tracts and secretory glands—constitute an anatomical barrier that is highly effective in protecting against infection. These epithelial layers produce a variety of protective substances, including acidic pH, enzymes, binding proteins, and antimicrobial proteins and peptides. Phagocytosis—engulfment and internalization of particulate materials such as microbes—is mediated by receptors on phagocytes (monocytes, macrophages, neutrophils, and dendritic cells) that either directly recognize PAMPs on the surface of microbes or recognize soluble proteins (opsonins) that bind to the microbes. PAMP binding triggers microbe uptake into phagosomes, which fuse with lysosomes or prepackaged granules, leading to their killing through enzymatic degradation, antimicrobial proteins and peptides, and toxic effects of reactive oxygen species (ROS) and reactive nitrogen species (RNS). Families of PRRs (TLRs, CLRs, RLRs, and NLRs) recognize a wide variety of PAMP (and DAMP) ligands and trigger signaling pathways that activate genes encoding proteins that contribute to innate and inflammatory responses. Vertebrate Toll-like receptors (TLRs, homologous to the fruit fly Toll receptor) are dimers of chains with extracellular leucine-rich (LRR) domains that bind PAMPs and DAMPs. Of the 13 TLRs found in mice and humans, some are on the cell surface, while others are in endosomes and lysosomes, where each binds PAMPs revealed by the disassembly or degradation of pathogens, such as microbial nucleic acids. TLR binding of PAMPs activates signaling pathways; the particular pathway varies depending on the TLR and the adaptor protein that binds to its cytoplasmic TIR domain (either MyD88 or TRIF). The signaling pathways activate transcription factors NF-B, various interferon regulating factors (IRFs), and transcription factors (such as AP-1) downstream of MAP kinase (MAPK) pathways. ■ ■ ■ ■ ■ ■ C-type lectin receptors (CLRs) comprise a heterogeneous family of cell surface PRRs that recognize cell wall components, largely sugars and polysaccharides, of bacteria and fungi. They trigger a variety of distinct signaling pathways that activate transcription factors, of which some are similar to those activated by TLRs. RIG-I-like receptors (RLRs) are RNA helicases that function as cytosolic PRRs. Most PAMPs that they recognize are viral double-stranded RNAs; after PAMP binding, RLRs trigger signaling pathways that activate IRFs and NF-B. NOD-like receptors (NLRs) are a large family of cytosolic PRRs activated by intracellular PAMPs, DAMPs, and other harmful substances. The NOD NLRs bind intracellular microbial components such as cell wall fragments. Other NLRs, such as NLRP3, function as inflammasomes; they are activated by a wide variety of PAMPs, DAMPs, and crystals by sensing changes in the intracellular milieu. Inflammasomes then activate the caspase-1 protease, which cleaves the inactive large precursors of the proinflammatory cytokines IL-1 and IL-18, so that active cytokines can be released from cells. The signaling pathways downstream of PRRs activate expression of a variety of genes, including those for antimicrobial peptides, Type 1 interferons (potent antiviral agents), cytokines (including proinflammatory IL-1, TNF-, and IL-6), chemokines, and enzymes that help to generate antimicrobial and inflammatory responses. Inflammatory responses are activated by the innate immune response to local infection or tissue damage—in particular, by the proinflammatory cytokines and certain chemokines that are produced. Key early components of inflammatory responses are increased vascular permeability, allowing soluble innate mediators to reach the infected or damaged site, and the recruitment of neutrophils and other leukocytes from the blood into the site. A later component of inflammatory responses is the acute phase response (APR), induced by certain proinflammatory cytokines (IL-1, TNF-, and IL-6). The APR involves increased synthesis and secretion by the liver of several antimicrobial proteins, including MBL, CRP, and complement components, which activate a variety of processes that contribute to eliminating pathogens. Innate Immunity ■ ■ ■ ■ ■ Natural killer (NK) cells are lymphocytes with innate immune functions. They express a set of activating receptors that recognize surface components of the body’s cells induced by infection, malignant transformation, or other stresses. Activated NK cells can kill the altered self cell and/or produce cytokines that help to induce adaptive immune responses to the altered cell. The importance of innate and inflammatory responses is demonstrated by the impact of a variety of genetic defects in humans. Defects in PRRs and signaling pathways activating innate responses lead to increased susceptibility to certain infections, while other defects that constitutively activate inflammasomes contribute to a variety of inflammatory disorders. As innate and inflammatory responses can be harmful as well as helpful, they are highly regulated by positive and negative feedback pathways. To escape elimination by innate immune responses, pathogens have evolved a wide array of strategies to block antimicrobial responses. There are many interactions between the innate and adaptive immune systems. The adaptive immune system has | CHAPTER 5 181 co-opted several pathogen-clearance mechanisms, such as opsonization and complement activation, so that they contribute to antibody-mediated pathogen elimination. ■ ■ ■ B cells and some T cells express TLRs, which can serve as costimulatory receptors. PAMP binding helps activate these cells to generate adaptive immune responses. TLRs expressed by NK cells can serve as activating receptors. Dendritic cells are a key cellular bridge between innate and adaptive immunity. Microbial components acquired by dendritic cells via PRR binding during the innate response are brought from the site of infection to lymph nodes. Activation of a dendritic cell by PAMPs stimulates the cell to mature, so that it acquires the abilities to activate naïve T cells to become mature cytotoxic and helper T cells and to influence the sets of cytokines that the mature TH will produce. Innate immunity appeared early during the evolution of multicellular organisms. While TLRs are unique to animals, PRRs with leucine-rich repeat (LRR) ligand-binding domains that induce innate immune responses are found in virtually all plants and animals. R E F E R E N C E S Areschoug, T., and S. Gordon. 2009. Scavenger receptors: role in innate immunity and microbial pathogenesis. Cellular Microbiology 11:1160. Fukata, M., et al. 2009. Toll-like receptors (TLRs) and Nod-like receptors (NLRs) in inflammatory disorders. Seminars in Immunology 21:242. Beutler, B., and E. T. Rietschel. 2003. Innate immune sensing and its roots: The story of endotoxin. Nature Reviews Immunology 3:169. Geijtenbeek, T. B. H., and S. I. Gringhuis. 2009. Signaling through C-type lectin receptors: Shaping immune repertoires. Nature Reviews Immunology 9:465. Bowie, A. G., and L. Unterholzner. 2008. Viral evasion and subversion of pattern-recognition receptor signalling. Nature Reviews Immunology 8:911. Grimsley, C., and K. S. Ravichandran. 2003. Cues for apoptotic cell engulfment: Eat-me, don’t-eat-me, and come-get-me signals. Trends in Cellular Biology 13:648. Bulet, P., et al. 2004. Anti-microbial peptides: from invertebrates to vertebrates. Immunology Review 198:169–184. Herrin, B. R., and M. D. Cooper. 2010. Alternative adaptive immunity in jawless vertebrates. Journal of Immunology 185:1367. Chroneos, Z. C., et al. 2010. Pulmonary surfactant: An immunological perspective. Cellular Physiology and Biochemistry 25:13. Coffman, R., et al. 2010. Vaccine adjuvants: Putting innate immunity to work. Immunity 33:492–503. Davis, B. K., et al. 2011. The inflammasome NLRs in immunity, inflammation, and associated diseases. Annual Review of Immunology 11:707. deFranco, A., et al. 2007. Immunity (Primers in Biology). Sunderland, MA: Sinauer Associates. Elinav, E., et al. 2011. Regulation of the antimicrobial response by NLR proteins. Immunity 34:665. Froy, O. 2005. Regulation of mammalian defensin expression by Toll-like receptor-dependent and independent signalling pathways. Cellular Microbiology 7:1387. Kawai, T., and S. Akira. 2010. The role of pattern recognition receptors in innate immunity: Update on Toll-like receptors. Nature Immunology 11:373. Kerrigan, A. M., and G. D. Brown. 2009. C-type lectins and phagocytosis. Immunobiology 214:562. Lemaitre, B. 2004. The road to Toll. Nature Reviews Immunology 4:521. Litvack, M. L., and N. Palaniyar. 2010. Soluble innate immune pattern-recognition proteins for clearing dying cells and cellular components: Implications on exacerbating or resolving inflammation. Innate Immunity 16:191. Loo, Y-M., and M. Gale. 2011. Immune signaling by RIG-I-like receptors. Immunity 34:680. Medzhitov, R., et al. 1997. A human homologue of the Toll protein signals activation of adaptive immunity. Nature 388:394. 182 PA R T I I | Innate Immunity Nathan, C., and A. Ding. 2010. SnapShot: Reactive oxygen intermediates (ROI). Cell 140:952. Palm, N. W., and R. Medzhitov. 2009. Pattern recognition receptors and control of adaptive immunity. Immunology Review 227:221. Poltorak, A., et al. 1998. Defective LPS signaling in C3Hej and C57BL/10ScCr mice: Mutations in TLR4 gene. Science 282:2085. Salzman, N. H., et al. 2010. Enteric defensins are essential regulators of intestinal microbial ecology. Nature Immunology 11:76. Schroder, K., and J. Tschopp. 2010. The inflammasome. Cell 140:821. Sun, J. C., et al. 2011. NK cells and immune “memory.” Journal of Immunology 186:1891. Takeuchi, O., and S. Akira. 2010. Pattern recognition receptors and inflammation. Cell 140:805. Tecle, T., et al. 2010. Defensins and cathelicidins in lung immunity. Innate Immunity 16:151. Tieu, D. T., et al. 2009. Alterations in epithelial barrier function and host defense responses in chronic rhinosinusitis. Journal of Clinical Immunology 124:37. van de Vosse, E., et al. 2009. Genetic deficiencies of innate immune signaling in human infectious disease. The Lancet Infectious Diseases 9:688. Willingham, S. B., et al. 2012. The CD47-signal regulatory protein alpha (SIRP1a) interaction is a therapeutic target for human solid tumors. Proceedings of the National Academy of Sciences USA 109:6662. Yu, M., and S. J. Levine. 2011. Toll-like receptor 3, RIG-I-like receptors and the NLRP3 inflammasome: Key modulators of innate immune responses to double-stranded viruses. Cytokine & Growth Factor Reviews 22:63. Useful Web Sites cpmcnet.columbia.edu/dept/curric-pathology/ pathology/pathology/pathoatlas/GP_I_menu. html Images are shown of the major inflammatory cells involved in acute and chronic inflammation, as well as examples of specific inflammatory diseases. portal.systemsimmunology.org/portal/web/guest/ homepage Systems Approach to Immunology (systemsimmunology.org) is a large collaborative research program formed to study the mechanisms by which the immune system responds to infectious disease by inciting innate inflammatory reactions and instructing adaptive immune responses. www.biomedcentral.com/1471-2172/9/7 Web site of The Innate Immunity Database, an NIH-funded multiinstitutional project that assembled microarray data on expression levels of more than 200 genes in macrophages stimulated with a panel of TLR ligands. The database is intended to support systems biology studies of innate responses to pathogens. www.immgen.org/index_content.html The Immunological Genome Project is a new cooperative effort for deep transcriptional profiling of all immune cell types. www.ncbi.nlm.nih.gov/PubMed PubMed, the National Library of Medicine database of more than 15 million publications, is the world’s most comprehensive bibliographic database for biological and biomedical literature. It is also highly user friendly, searchable by general or specific topics, authors, reviews, and so on. It is the best resource to use for finding the latest research articles on innate immunity or other topics in the biomedical sciences. animaldiversity.ummz.umich.edu/site/index. html The Animal Diversity Web (ADW), based at the University of Michigan, is an excellent and comprehensive database of animal classification as well as a source of information on animal natural history and distribution. Here is a place to find information about animals that are not humans or mice. www.biolegend.com/media_assets/.../InnateImmunityResourceGuide_3.pdf Presents a concise sum- en.wikipedia.org/wiki/Innate_immune_system The mary of innate immunity, including a detailed table on the cells of the innate immune system, including their locations in the body, functions, products, receptors, and modes of activation. Wikipedia Web site presents a detailed summary of the innate immune system in animals and plants; it contains numerous figures and photographs illustrating various aspects of innate immunity, plus links to many references. S T U D Y Q U E S T I O N S CLINICAL FOCUS QUESTION What infections are unusually prevalent in individuals with genetic defects in TLRs or the MyD88-dependent TLR signaling pathway? in individuals with defects in pathways activating the production or antiviral activities of IFN-,? Why is it thought that these individuals aren’t susceptible to a wider range of diseases, and what evidence supports this hypothesis? 1. Use the following list to complete the statements that fol- low. Some terms may be used more than once or not at all. Innate Immunity Acute phase response NK cells Antibodies NLRs Arginine NO Caspase-1 O2 C-reactive protein (CRP) PAMPs Complement Phagocytosis Costimulatory molecules Proinflammatory cytokines Defensins PRRs Dendritic cells Psoriasin Ficolins RLRs IL-1 ROS Inflammasomes RNS iNOS Surfactant proteins (SP) A, D Interferons-, TRIF IRFs T-cell receptors Lysozyme TLR2 Mannose-binding lectin (MBL) TLR3 MyD88 TLR4 NADPH TLR7 NADPH phagosome oxidase TLR9 NF-B TNF- a. Examples of proteins and peptides with direct antimi- b. c. d. e. f. g. crobial activity that are present on epithelial surfaces are _____, _____, and _____. Soluble pattern-recognition proteins that function as opsonins, which enhance ____, include _____, _____, _____, and _____; of these, the ones that also activate complement are _____ and _____. The enzyme _____ uses _____ to generate microbekilling _____; one of these, plus the antimicrobial gas ____ generated by the _____ enzyme from the amino acid _____, are used to generate ____, which are also antimicrobial. As components of innate immune responses, both _____ (secreted proteins) and _____ (a type of lymphocyte) defend against viral infection. The _____ occurs when _____, such as _____ and _____, are generated by innate immune responses and act on the liver. _____, receptors of innate immunity that detect _____, are encoded by germline genes, whereas the signature receptors of adaptive immunity, _____ and _____, are encoded by genes that require gene rearrangements during lymphocyte development to be expressed. Among cell surface TLRs, _____ detects Gram-positive bacterial infections while _____ detects Gram-negative infections. | CHAPTER 5 183 h. Some cells use the intracellular TLRs _____ and _____ to detect RNA virus infections and _____ to detect infections by bacteria and some DNA viruses. i. _____ is unique among PRRs in that it functions both on the plasma membrane and in endosomes and binds both the _____ and _____ adaptor proteins. j. _____ are receptors that detect cytosolic viral nucleic acids, while _____ include cytosolic receptors that detect intracellular bacterial cell wall components. k. Key transcription factors for inducing expression of proteins involved in innate immune responses are _____ and _____. l. The production of the key proinflammatory cytokine _____ is complex, as it requires transcriptional activation by signaling pathways downstream of _____ followed by cleavage of its large precursor protein by _____, which is activated by _____, members of the _____ family of innate receptors. m. After maturation induced by binding of _____ to their _____, cells known as _____ become efficient activators of naïve helper and cytotoxic T cells. n. _____ and _____ are innate immune system components common to both plants and animals. 2. What were the two experimental observations that first linked TLRs to innate immunity in vertebrates? 3. What are the hallmark characteristics of a localized inflam- matory response? How are they induced by the early innate immune response at the site of infection, and how do these characteristics contribute to an effective innate immune response? 4. In vertebrates, innate immunity collaborates with adaptive immunity to protect the host. Discuss this collaboration, naming key points of interaction between the two systems. Include at least one example in which the adaptive immune response contributes to enhanced innate immunity. 5. As adaptive immunity evolved in vertebrates, the more ancient system of innate immunity was retained. Can you think of any disadvantages to having a dual system of immunity? Would you argue that either system is more essential? ANALYZE THE DATA A variety of studies have shown that hel- minth worm parasites have developed numerous mechanisms for inhibiting innate and inflammatory responses that otherwise would contribute to their elimination. One particular parasite, the filarial nematode Acanthocheilonema viteae, has been shown to secrete a lipoglycoprotein, ES-62, that inhibits proinflammatory responses. Based on these initial observations, a group of investigators has been studying how ES-62 achieves this anti-inflammatory effect and whether ES-62 might inhibit the development of septic shock in individuals with sepsis. Interestingly, they found that ES-62 binds to TLR4 (but not to any other TLR). Based on what you have learned in this chapter and the information and data provided below from these investigators’ recent paper (P. Puneet et al., 2011, Nature Immunology 12:344), answer the following questions. (Note: The various methods used to obtain the data below are described in Chapter 20.) PA R T I I 184 | Innate Immunity Basal LPS LPS + ES-62 BLP BLP + ES-62 Mediator release (ng/ml) 2.8 2.4 2.0 1.6 1.2 0.8 0.4 0 * * * IL-1β IL-6 * * * * * IL-8 * TNF HMGB1 (a) Effects of ES-62 on proinflammatory mediator production Macrophages 100 100 Events Events 150 50 101 102 103 50 0 0 10 104 Macrophages RANTES 101 102 103 104 Neutrophils 100 Events Events * TLR4 100 50 0 0 10 MCP-1 * Isotype Normal Sepsis Sepsis + ES-62 TLR4 (b) * Neutrophils 150 0 0 10 * flammatory cytokines IL-1, IL-6, and TNF-; the chemokines IL-8 (CXCL8), MCP-1 (CCL2), and RANTES (CCL5); and the nuclear protein HMGB1 (which is released by activated or damaged cells and has proinflammatory activities). Macrophages from patients with sepsis were incubated with nothing (“Basal”), or with the TLR4 ligand LPS or the TLR2 lipoprotein ligand BLP with or without ES-62. After 24 hours, tissue culture fluids were harvested and assayed by ELISA for levels of the proin(a) ** 101 102 103 104 TLR2 (b) Effects of ES-62 on cell surface expression of TLR4 and TLR2 Levels of TLR4 and TLR2 were measured on macrophages and neutrophils from normal people and from patients with sepsis incubated with or without ES-62 for 3 hours. Cells were stained with fluorescent antibodies to TLR4 or TLR2, or with control (nonreactive) antibodies Isotype Normal Sepsis Sepsis + ES-62 50 0 0 10 101 102 103 104 TLR2 with the same immunoglobulin heavy-chain class (“Isotype”). The stained cells were analyzed using a flow cytometer for the levels of their fluorescence, indicated on the X-axis. The fluorescence obtained with the Isotype control antibody represents no specific staining. The number of cells (“events”) with different levels of fluorescence is indicated on the Y-axis. (c) Effect of ES-62 on MyD88 levels Cell extracts were prepared from macrophages and neutrophils from patients with sepsis that were incubated without (0) or for 3 or 6 hours with ES-62. The extract proteins were separated by size by SDS-PAGE and subjected to Western blots to assess levels of MyD88 and, as a control, the ubiquitous cellular protein -tubulin. The gels were blotted onto filter paper and the filters stained with antibodies to MyD88 and to -tubulin. Bound antibodies were visualized to reveal the levels of MyD88 and tubulin. Innate Immunity a. What is septic shock, and how is it induced by bacterial infections? b. The investigators tested whether ES-62 inhibits the production of proinflammatory mediators by macrophages from normal people and from patients hospitalized with sepsis. Shown in panel (a) above are the results with macrophages from patients with sepsis; cells from normal individuals gave similar results. What effect does ES-62 have on the responses induced by the two PAMPs? Given that ES-62 binds to TLR4 but not TLR2, what do these results say about the mechanism of ES-62 inhibition? The investigators compared the levels of TLR4 and TLR2 expressed by macrophages and neutrophils from normal people with those from patients with sepsis. As had been seen by others, panel (b) shows that cells from patients with sepsis express higher levels of these TLRs than do cells from normal individuals. Of what kind of feedback regulation is this an example? Why is this feedback response normally beneficial in individuals with limited infections but disadvantageous in individuals with sepsis? | CHAPTER 5 185 In the same experiment shown in panel (b), some cells from patients with sepsis were incubated with ES-62. What effect does ES-62 have on TLR4 expression? on TLR2 expression? The investigators also looked at the levels of MyD88 in macrophages and neutrophils from sepsis patients after no exposure or 3 or 6 hours of exposure to ES-62 (panel [c]). What does this Western blot show? Returning to the initial finding that ES-62 inhibits the induction of proinflammatory mediators by PAMPs that activate either TLR4 or TLR2 [panel (a)], how might that finding be explained by the data shown in panels (b) and (c) (i.e., the effects of ES-62 on TLR and MyD88 expression)? Given these results, do you think injection of ES-62 might be used to cure septic shock once it has occurred? Do you think ES-62 might prevent septic shock from occurring in patients with sepsis? Explain your answers. This page lelt intentionally blank. 6 The Complement System T he term complement (spelled with an e) refers to a set of serum proteins that cooperates with both the innate and the adaptive immune systems to eliminate blood and tissue pathogens. Like the components of the blood clotting system, complement proteins interact with one another in catalytic cascades. As mentioned in Chapters 3 and 5, various complement components bind and opsonize bacteria, rendering them susceptible to receptor-mediated phagocytosis by macrophages, which express membrane receptors for complement proteins. Other complement proteins elicit inflammatory responses, interface with components of the adaptive immune system, clear immune complexes from the serum, and/or eliminate apoptotic cells. Finally, a Membrane Attack Complex (MAC) assembled from complement proteins directly kills some pathogens by creating pores in microbial membranes. The biological importance of complement is emphasized both by the pathological consequences of mutations in the genes encoding complement proteins as well as by the broad range of strategies that have evolved in microorganisms to evade it. Research on complement began in the 1890s when Jules Bordet at the Institut Pasteur in Paris showed that sheep antiserum to the bacterium Vibrio cholerae caused lysis (membrane destruction) of the bacteria, and that heating the antiserum destroyed its bacteriolytic activity. Surprisingly, the ability to lyse the bacteria was restored to the heated serum by adding fresh serum that contained no antibacterial antibodies. This finding led Bordet to reason that bacteriolysis required two different substances: the heat-stable specific antibodies that bound to the bacterial surface, and a second, heat-labile (sensitive) component responsible for the lytic activity. In an effort to purify this second, nonspecific component, Bordet developed antibodies specific for red blood cells and used these, along with purified serum fractions, to identify those fractions that cooperated with the antibodies to induce lysis of the red blood cells (hemolysis). The famous immunologist Paul Ehrlich, working independently in Berlin, carried out similar Sheep red blood cells (red) are phagocytosed by macrophages (green) after opsonization by complement. ■ The Major Pathways of Complement Activation ■ The Diverse Functions of Complement ■ The Regulation of Complement Activity ■ Complement Deficiencies ■ Microbial Complement Evasion Strategies ■ The Evolutionary Origins of the Complement System experiments and coined the term complement, defining it as “the activity of blood serum that completes the action of antibody.” In the ensuing years, researchers have discovered that the action of complement is the result of interactions among a complex group of more than 30 glycoproteins. Most complement components are synthesized in the liver by hepatocytes, although some are also produced by other cell types, including blood monocytes, tissue macrophages, fibroblasts, and epithelial cells of the gastrointestinal and genitourinary tracts. Complement components constitute approximately 15% of the globulin protein fraction in plasma, and their combined concentration can be as high as 3 mg/ml. In addition, several of the regulatory components of the system exist on cell membranes, so the term complement therefore now embraces glycoproteins distributed among the blood plasma and cell membranes. Complement components can be classified into seven functional categories (Overview Figure 6-1): 187 188 PA R T I I | Innate Immunity 6-1 OVERVIEW FIGURE Proteins Involved in the Complement System 1 2 Convertase activators (C1r, C1s, C4b, C2a) and enzymatic mediators (C3 convertase, C5 convertase) Inititators (C1q, MBL, ficolins) + C5b C6 C7 Phagocyte 3 Opsonins 4 5 Anaphylatoxins C8C9 C9 9 C9 C9 C9 Membrane attack complex 6 Complement receptors Inflammation Degrade complement components 7 (1) The complement pathways are initiated by proteins that bind to pathogens, either directly or via an antibody or other pathogenspecific protein. After a conformational change, (2) enzymatic mediators activate other enzymes that generate the central proteins of the complement cascade, the C3 and C5 convertases, which cleave C3 and C5, releasing active components that mediate all functions of complement, including (3) opsonization, (4) inflammation, and (5) the generation of the membrane attack Prevent deposition of components Regulatory proteins complex (MAC). Effector complement proteins can label an antibody-antigen complex for phagocytosis (opsonins), initiate inflammation (anaphylatoxins), or bind to a pathogen and nucleate the formation of the MAC. Often, these effectors act through (6) complement receptors on phagocytic cells, granulocytes, or erythrocytes. (7) Regulatory proteins limit the effects of complement by promoting their degradation or preventing their binding to host cells. The Complement System 1. 2. 3. Initiator complement components. These proteins initiate their respective complement cascades by binding to particular soluble or membrane-bound molecules. Once bound to their activating ligand, they undergo conformational alterations resulting in changes in their biological activity. The C1q complex, Mannose Binding Lectin (MBL), and the ficolins are examples of initiator complement components. Enzymatic mediators. Several complement components, (e.g., C1r, C1s, MASP2, and factor B) are proteolytic enzymes that cleave and activate other members of the complement cascade. Some of these proteases are activated by binding to other macromolecules and undergoing a conformational change. Others are inactive until cleaved by another protease enzyme and are thus termed zymogens: proteins that are activated by proteolytic cleavage. The two enzyme complexes that cleave complement components C3 and C5, respectively, are called the C3 and C5 convertases and occupy places of central importance in the complement cascades. Membrane-binding components or opsonins. Upon activation of the complement cascade, several proteins are cleaved into two fragments, each of which then takes on a particular role. For C3 and C4, the larger fragments, C3b and C4b, serve as opsonins, enhancing phagocytosis by binding to microbial cells and serving as binding tags for phagocytic cells bearing receptors for C3b or C4b. As a general rule, the larger fragment of a cleaved complement component is designated with the suffix “b,” and the smaller with the suffix “a.” However, there is one exception to this rule: the larger, enzymatically active form of the C2 component is named C2a. 4. Inflammatory mediators. Some small complement fragments act as inflammatory mediators. These fragments enhance the blood supply to the area in which they are released, by binding to receptors on endothelial cells lining the small blood vessels and inducing an increase in capillary diameter. They also attract other cells to the site of tissue damage. Because such effects can be harmful in excess, these fragments are called anaphylatoxins, meaning substances that cause anaphylaxis (“against protection”). Examples include C3a, C5a, and C4a. 5. Membrane attack proteins. The proteins of the membrane attack complex (MAC) insert into the cell membranes of invading microorganisms and punch holes that result in lysis of the pathogen. The complement components of the MAC are C5b, C6, C7, C8, and multiple copies of C9. 6. Complement receptor proteins. Receptor molecules on cell surfaces bind complement proteins and signal specific cell functions. For example, some complement receptors such as CR1 bind to complement components such as C3b on the surface of pathogens, triggering phagocytosis of the C3-bound pathogen. Binding of the complement component C5a to C5aR receptors on neutrophils stimulates | CHAPTER 6 189 neutrophil degranulation and inflammation. Complement receptors are named with “R,” such as CR1, CR2, and C5aR. 7. Regulatory complement components. Host cells are protected from unintended complement-mediated lysis by the presence of membrane-bound as well as soluble regulatory proteins. These regulatory proteins include factor I, which degrades C3b, and Protectin, which inhibits the formation of the MAC on host cells. This chapter describes the components of the complement system, their activation via three major pathways, the effector functions of the molecules of the complement cascade, as well as their interactions with other cellular and molecular components of innate and adaptive immunity. In addition, it addresses the mechanisms that regulate the activity of these complement components, the evasive strategies evolved by pathogens to avoid destruction by complement, and the evolution of the various complement proteins. This chapter’s Classic Experiment tells the story of the scientist who discovered the alternative pathway of complement. In the Clinical Focus segment, we address various therapies that target elements of the complement cascades. Finally, an Advances Box describes some of the many tactics used by Staphylococcus aureus bacteria to escape complement-mediated destruction. The Major Pathways of Complement Activation Complement components represent some of the most evolutionarily ancient participants in the immune response. As viruses, parasites, and bacteria have attacked vertebrate hosts and learned to evade one aspect of the complement system, alternative pathways have evolved in an endless dance of microbial attack and host response. In the complex pathways that follow, we can gain a glimpse of the current state of what is actually an ongoing evolutionary struggle on the part of host organisms to combat microbial infection, while minimizing damage to their own cells. The major initiation pathways of the complement cascade are shown in Figure 6-2. Although the initiating event of each of the three pathways of complement activation is different, they all converge in the generation of an enzyme complex capable of cleaving the C3 molecule into two fragments, C3a and C3b. The enzymes that accomplish this biochemical transformation are referred to as C3 convertases. As illustrated in Figure 6-2, the classical and lectin pathways use the dimer C4b2a for their C3 convertase activity, while the alternative pathway uses C3bBb to achieve the same end; however, the final result is the same: a dramatic increase in the concentration of C3b, a centrally located and multifunctional complement protein. The second set of convertase enzymes of the cascade, C5 convertases, are formed by the addition of a C3b component 190 PA R T I I | Innate Immunity Classical pathway Lectin pathway Alternative pathway Antigen-antibody immune complexes PAMP recognition by lectins Spontaneous hydrolysis or pathogenic surfaces Initiation MBL C1q C3(H2O)Bb C3 C3 Antibody Carbohydrates C3a MASPs C1r2s2 C3b C3bBb Factors B, D, Properdin C2 C4 Amplification C4a C3 convertases C5 convertases C3bBb C3bBbC3b C4b2a C4b2a3b C3a C3b C5 C3 Termination Inflammation FIGURE 6-2 The generation of C3 and C5 convertases by the three major pathways of complement activation. The classical pathway is initiated when C1q binds to antigen-antibody complexes. The antigen is shown here in dark red and the initiating antibody in green. The C1r enzymatic component of C1 (shown in blue) is then activated and cleaves C1s, which in turn cleaves C4 to C4a (an anaphylatoxin, bright red) and C4b. C4b attaches to the membrane, and binds C2, which is then cleaved by C1s to form C2a and C2b. (C2b is then acted upon further to become an inflammatory mediator.) C2a remains attached to C4b, forming the classical pathway C3 convertase (C4b2a). In the lectin pathway, mannosebinding lectin (MBL, green) binds specifically to conserved carbohy- to the C3 convertases. C5 convertases cleave C5 into the inflammatory mediator, C5a, and C5b, which is the initiating factor of the MAC. We will now describe each of these pathways in more detail. The proteins involved in each of these pathways are listed in Table 6-1. The Classical Pathway Is Initiated by Antibody Binding The classical pathway of complement activation is considered part of the adaptive immune response since it begins with the formation of antigen-antibody complexes. These complexes may be soluble, or they may be formed when an antibody binds to antigenic determinants, or epitopes, situated on viral, fungal, parasitic, or bacterial cell membranes. Soluble Opsonization C5b C5a Lysis Inflammation drate arrays on pathogens, activating the MBL-associated serine proteases (MASPs, blue). The MASPs cleave C2 and C4 generating the C3 convertase as in the classical pathway. In the alternative pathway, C3 undergoes spontaneous hydrolysis to C3(H2O), which binds serum factor B. On binding to C3(H2O), B is cleaved by serum factor D, and the resultant C3(H2O)Bb complex forms a fluid phase C3 convertase. Some C3b, released after C3 cleavage by this complex, binds to microbial surfaces. There, it binds factor B, which is cleaved by factor D, forming the cell-bound alternative pathway C3 convertase, C3bBb. This complex is stabilized by properdin. The C5 convertases are formed by the addition of a C3b fragment to each of the C3 convertases. antibody-antigen complexes are often referred to as immune complexes, and only complexes formed by IgM or certain subclasses of IgG antibodies are capable of activating the classical complement pathway (see Chapter 13). The initial stage of activation involves the complement components C1, C2, C3, and C4, which are present in plasma as zymogens. The formation of an antigen-antibody complex induces conformational changes in the nonantigen-binding (Fc) portion of the antibody molecule. This conformational change exposes a binding site for the C1 component of complement. In serum, C1 exists as a macromolecular complex consisting of one molecule of C1q and two molecules each of the serine proteases, C1r and C1s, held together in a Ca2⫹-stabilized complex (C1qr2s2) (Figure 6-3a). The C1q molecule itself is composed of 18 polypeptide chains that associate to form six collagen-like triple helical arms, the tips The Complement System TABLE 6-1 | CHAPTER 6 191 Proteins of the three major pathways of complement activation Molecule Biologically active fragments Biological function Active in which pathway IgM, IgG Binding to pathogen surface and initiating complement cascade Classical pathway Mannose-binding lectin, or ficolins Binding to carbohydrates on microbial surface and initiating complement cascade Lectin pathway Initiation of the classical pathway by binding Ig Binding to apoptotic blebs and initiating phagocytosis of apoptotic cells Serine protease, cleaving C1r and C1s Serine protease, cleaving C4 and C2 Classical pathway MASP-1 MBL-Associated Serine Protease 1. MASP-2 appears to be functionally the more relevant MASP-protease. Lectin pathway MASP-2 Serine protease. In complex with MBL/ficolin and MASP-2, cleaves C4 and C2. Lectin pathway C2a* Serine protease. With C1 and C4b, is a C3 convertase. Classical and lectin pathways C2b* Inactive in complement pathway. Cleavage of C2b by plasmin releases C2 kinin, a peptide that stimulates vasodilation. C4b Binds microbial cell membrane via thioester bond. With C1 and C2a, is a C3 convertase. Has weak anaphylatoxin (inflammatory) activity Proteolytic cleavage products generated by factor I C1 C1q (C1r)2 (C1s)2 C2 C4 C4a C4c, C4d C3 C3a C3b C3(H2O) iC3b and C3f C3c and C3d or C3dg Factor B Ba Bb Factor D Anaphylatoxin. Mediates inflammatory signals via C3aR. Potent in opsonization, tagging immune complexes, pathogens, and apoptotic cells for phagocytosis With C4b and C2a, forms the C5 convertase With Bb, forms the C3 convertase With Bb and one more molecule of C3b (C3bBb3b) acts as a C5 convertase C3 molecule in which the internal thioester bond has undergone hydrolysis With Bb, acts as a fluid phase C3 convertase. Proteolytic fragments of C3b, generated by factor I. iC3b binds receptors CR3, CR4, and CRIg; CR2 binds weakly. Proteolytic fragments of iC3b generated by factor I. C3d/dg bind antigen and to CR2, facilitating antigen-binding to B cells. C3c binds CRIg on fixed tissue macrophages Binds C3(H2O) and is then cleaved by factor D into two fragments: Ba and Bb Smaller fragment of factor D-mediated cleavage of factor B. May inhibit proliferation of activated B cells. Larger fragment of factor D-mediated cleavage of factor B With C3(H2O), acts as fluid phase C3 convertase With C3b, acts as cell-bound C3 convertase With two molecules of C3b, acts as C5 convertase Proteolytic enzyme that cleaves factor B into Ba and Bb only when it is bound to either C3(H2O) or to C3b Classical and lectin pathways Classical pathway Tickover and properdin alternative pathways Tickover and properdin alternative pathways Alternative pathway Alternative pathway Alternative pathway Alternative pathway Alternative pathway Alternative pathway (continued) 192 TABLE 6-1 PA R T I I | Innate Immunity (continued) Factor P (properdin) C5 Stabilizes the C3bBb complex on microbial cell surface C5a C5b Alternative pathway Anaphylatoxin binding to C5aR induces inflammation Component of Membrane Attack Complex (MAC). Binds cell membrane and facilitates binding of other components of the MAC. All C6 Component of MAC. Stabilizes C5b. In the absence of C6, C5b is rapidly degraded. All C7 Component of MAC. Binds C5bC6 and induces conformational change allowing C7 to insert into interior of membrane. All C8 Component of MAC. Binds C5bC6C7 and creates a small pore in membrane. All C9 Component of MAC. 10–19 molecules of C9 bind C5bC6C7C8 and create large pore in membrane. All * Because C2a is the larger, active fragment of C2, some writers have tried to alter the nomenclature in order to make C2 conform to the convention that the “b” fragment is the larger, active fragment and the “a” fragment is smaller and may be an anaphylatoxin. However, this effort does not appear to be making headway. Note that the smaller fragment of C2, which we name C2b, is inactive in the complement pathway. (a) (b) C1q 2 C1 complex 3 C1r2s2 s r IgG antibody (c) Collagen-like triple helix Binds CH2 domain of antibody FIGURE 6-3 Structure of the C1 macromolecular complex. (a) C1q interacts with two molecules each of C1r and C1s to create the C1 complex. (b) The C1q molecule consists of 18 polypeptide chains in six collagen-like triple helices, each of which contains one A, B, and C chain, shown in three different shades of green. The head group of each triplet contains elements of all three chains. (c) Electron micrograph of C1q molecule showing stalk and six globular heads. The dark circle is a gold-labeled molecule of fibromodulin, a component of cartilage that binds the head groups of C1q. [Sjoberg, A., et al. September 16, 2005 The J. Biological Chem, Vol. 280, Issue 37, 3230132308. ?2005 by the Merican Society for Biochemistry and Molecular Biology.] The Complement System | CHAPTER 6 193 (a) (b) FIGURE 6-4 Models of pentameric IgM in planar form (a) and “staple” form (b). Several C1q-binding sites in the Fc region are accessible in the staple form, whereas none are exposed in the planar form. [From A. Feinstein et al., 1981, Monographs in Allergy 17:28, and 1981, Annals of the New York Academy of Sciences 190:1104.] of which bind the CH2 domain of the antigen-bound antibody molecule (Figure 6-3b, c). Each C1 macromolecular complex must bind by its C1q globular heads to at least two Fc sites for a stable C1-antibody interaction to occur. Recall from Chapter 3 that serum IgM exists as a pentamer of the basic four-chain immunoglobulin structure. Circulating, nonantigen-bound IgM adopts a planar configuration (Figure 6-4a), in which the C1qbinding sites are not exposed. However, when pentameric IgM is bound to a multivalent antigen, it undergoes a conformational change, assuming the so-called “staple” configuration (Figure 6-4b), in which at least three binding sites for C1q are exposed. Thus, an IgM molecule engaged in an antibody-antigen complex can bind C1q, whereas circulating, nonantigen-bound IgM cannot. In contrast to pentameric IgM, monomeric IgG contains only one C1q binding site per molecule, and the conformational changes IgG undergoes on antigen binding are much more subtle than those experienced by IgM. There is therefore a striking difference in the efficiency with which IgM and IgG are able to activate complement. Less than 10 molecules of IgM bound to a red blood cell can activate the classical complement pathway and induce lysis, whereas some 1000 molecules of cell-bound IgG are required to ensure the same result. The intermediates in the classical activation pathway are depicted schematically in Figure 6-5. Proteins of the classical pathway are numbered in the order in which the proteins were discovered, which does not quite correspond with the order in which the proteins act in the pathway (a disconnect that has troubled generations of immunology students). Note that binding of one component to the next always induces either a conformational change or an enzymatic cleavage, which allows for the next reaction in the sequence to occur. Binding of C1q to the CH2 domains of the Fc regions of the antigen-complexed antibody molecule induces a conformational change in one of the C1r molecules that converts it to an active serine protease enzyme. This C1r molecule then cleaves and activates its partner C1r molecule. The two C1r proteases then cleave and activate the two C1s molecules (see Figure 6-5, part 1). C1s has two substrates, C4 and C2. C4 is activated when C1s hydrolyzes a small fragment (C4a) from the amino terminus of one of its chains (see Figure 6-5, part 2). The C4b fragment attaches covalently to the target membrane surface in the vicinity of C1, and then binds C2. C4b binding to the membrane occurs when an unstable, internal thioester on C4b, exposed upon C4 cleavage, reacts with hydroxyl or amino groups of proteins or carbohydrates on the cell membrane. This reaction must occur quickly, otherwise the thioester C4b is further hydrolyzed and can no longer make a covalent bond with the cell surface (Figure 6-6). Approximately 90% of C4b is hydrolyzed before it can bind the cell surface. On binding C4b, C2 becomes susceptible to cleavage by the neighboring C1s enzyme, and the smaller C2b fragment diffuses away, leaving behind an enzymatically active C4b2a complex. In this complex, C2a is the enzymatically active fragment, but it is only active when bound by C4b. This C4b2a complex is called C3 convertase, referring to its role in converting C3 into an active form. The smaller fragment generated by C4 cleavage, C4a, is an anaphylatoxin, and its function is described below. The membrane-bound C3 convertase enzyme, C4b2a, now hydrolyzes C3, generating two unequal fragments; the small anaphylatoxin C3a and the pivotal fragment C3b. A single C3 convertase molecule can generate over 200 molecules of C3b, resulting in tremendous amplification at this step of the classical pathway. PA R T I I 194 | Innate Immunity 6-5 OVERVIEW FIGURE Intermediates in the Classical Pathway of Complement Activation up to the Formation of the C5 Convertase 1 C1q binds antigen-bound antibody, and induces a conformational change in one C1r molecule, activating it. This C1r then activates the second C1r and the two C1s molecules. 3 C3 convertase hydrolyzes many C3 molecules. Some combine with C3 convertase to form C5 convertase. C1qr2s2 C1q Antibody C1r2s2 Antibody binding sites on antigen (epitopes) FC + C3b C3 C4b2a3b C5 convertase C4b2a 2 C1s cleaves C4 and C2. C4 is cleaved first and C4b binds to the membrane close to C1. C4b binds C2 and exposes it to the action of C1s. C1s cleaves C2, creating the C3 convertase, C4b2a. 4 C3a The C3b component of C5 convertase binds C5, permitting C4b2a to cleave C5. C4 C4a C5 C5 convertase C5a C5b C4b C4bC2 C4b2a C3 convertase C2b Antigenic determinants are shown in dark red, initiating components (antibodies and C1q) are shown in green, active enzymes are shown in blue, and anaphylatoxins in bright red. The generation of C3b is centrally important to many of the actions of complement. Deficiencies of complement components that act prior to C3 cleavage leave the host extremely vulnerable to both infectious and autoimmune diseases, whereas deficiencies of components later in the pathway are generally of lesser consequence. This is because C3b acts in three important and different ways to protect the host. First, in a manner very similar to that of C4b, C3b binds covalently to microbial surfaces, providing a molecular “tag” and thereby allowing phagocytic cells with C3b receptors to engulf the tagged microbes. This process is called opsonization. Second, C3b molecules can attach to the Fc portions of antibodies participating in soluble antigen-antibody complexes. These C3b-tagged immune complexes are bound by C3b receptors on phagocytes or red blood cells, and are | The Complement System S S S S S S SH S C OH NH2 NH Cell membrane O S S S S SH S C S S S NH2 O 195 (c) Hydrolyzed C4b S S S S S S C S S O Thioester bond S (b) Bound C4b S (a) Unbound C4b CHAPTER 6 Cell membrane Cell membrane FIGURE 6-6 Binding of C4b to the microbial membrane surface occurs through a thioester bond via an exposed amino or hydroxyl group. (a) Both C3b and C4b contain highly reactive thioesters, which are subject to nucleophilic attack by hydroxyl or amino groups on cell membrane proteins and carbohy- drates. (b) Breakage of the thioester leads to the formation of covalent bonds between the membrane macromolecules and the complement components. (c) If this covalent bond formation does not occur quickly after generation of the C3b and C4b fragments, the thioester will be hydrolyzed to a nonreactive form. either phagocytosed, or conveyed to the liver where they are destroyed. Finally, some molecules of C3b bind the membranelocalized C4b2a enzyme to form the trimolecular, membranebound, C5 convertase complex C4b2a3b. The C3b component of this complex binds C5, and the complex then cleaves C5 into the two fragments: C5b and C5a (see Figure 6-5 part 4). C4b2a3b is therefore the C5 convertase of the classical pathway. This trio of tasks accomplished by the C3b molecule places it right at the center of complement attack pathways. As we will see, C5b goes on to form the MAC with C6, C7, C8, and C9. dimensional array. Thus, one can think of MBL as a classic pattern recognition receptor (see Chapter 5). Consistent with MBL’s place at the beginning of an important immune cascade, many individuals with low levels of MBL suffer from repeated, serious bacterial infections. MBL is constitutively expressed by the liver and, like C1q, which it structurally resembles, MBL belongs to the subclass of lectins known as collectins (see Chapter 5). More recently, The Lectin Pathway Is Initiated When Soluble Proteins Recognize Microbial Antigens The lectin pathway, like the classical pathway, proceeds through the activation of a C3 convertase composed of C4b and C2a. However, instead of relying on antibodies to recognize the microbial threat and to initiate the complement activation process, this pathway uses lectins—proteins that recognize specific carbohydrate components primarily found on microbial surfaces—as its specific receptor molecules (Figure 6-7; see also Figure 5-7a). Mannose-binding lectin (MBL), the first lectin demonstrated to be capable of initiating complement activation, binds close-knit arrays of mannose residues that are found on microbial surfaces such as those of Salmonella, Listeria, and Neisseria strains of bacteria; Cryptococcus neoformans and Candida albicans strains of fungi; and even the membranes of some viruses such as HIV-1 and respiratory syncytial virus. The complement pathway that it initiates is referred to as the lectin pathway of complement activation. Further characterization of the sugars recognized by MBL demonstrated that MBL also recognizes structures in addition to mannose, including N-acetyl glucosamine, D-glucose, and L-fucose. All those sugars, including mannose, present their associated hydroxyl groups in a particular three- MBL MASP-1 or MASP-2 Polysaccharide antigen FIGURE 6-7 Initiation of the lectin pathway relies on lectin receptor recognition of microbial cell surface carbohydrates. Lectin receptors, such as MBL, bind microbial cell surface carbohydrates. Once attached to the carbohydrates, they bind the MASP family serine proteases, which cleave C2 and C4 with the formation of a lectin-pathway C3 convertase. 196 PA R T I I | Innate Immunity the ficolins (see Chapter 5), another family of proteins structurally related to the collectins, have been recognized as additional initiators of the lectin pathway of complement activation. L-ficolin, H-ficolin, and M-ficolin each bind specific types of carbohydrates on microbial surfaces. We will use MBL as our example in subsequent paragraphs. MBL is associated in the serum with MBL-Associated Serine Proteases, or MASP proteins. Three MASP proteins— MASP1, MASP2, and MASP3—have been identified, but most studies of MASP function point to the MASP2 protein as being the most important actor in the next step of the MBL pathway. MASP-2 is structurally related to the serine protease C1s, and can cleave both C2 and C4 (see Figure 6-2), giving rise to the C3 convertase, C4b2a, that we first encountered in our discussion of the classical pathway. Thus, the lectin pathway utilizes all the same components as the classical pathway with the single exception of the C1 complex. The soluble lectin receptor replaces the antibody as the antigen-recognizing component, and MASP proteins take the place of C1r and C1s in cleaving and activating the C3 convertase. Once the C3 convertase is formed, the reactions of the lectin pathway are the same as for the classical pathway; the C5 convertase of the lectin pathway, like that of the classical pathway, is also C4b2a3b. binds another serum protein, factor B (Figure 6-8a). When bound to C3(H2O), factor B becomes susceptible to cleavage by a serum protease, factor D. Factor D cleaves B, releasing a smaller Ba subunit, which diffuses away and leaves a catalytically active Bb subunit that remains bound to C3(H2O). This C3(H2O)Bb complex, which at this point is still in the fluid phase (i.e., in the plasma, not bound to any cells), has C3 convertase activity. It rapidly cleaves many molecules of C3 into C3a and C3b (Figure 6-8b). This initiating C3 convertase is constantly being formed in plasma, and breaking down a few C3 molecules, but it is then just as rapidly degraded. However, if there is an infection present, some of (b) (a) C3 C3 Initiation of the alternative pathway of complement activation is independent of antibody-antigen interactions, and so this pathway, like the lectin pathway, is also considered to be part of the innate immune system. However, unlike the lectin pathway, the alternative pathway uses its own set of C3 and C5 convertases (see Figure 6-2). As we will see, the alternative pathway C3 convertase is made up of one molecule of C3b and one molecule unique to the alternative pathway, Bb. A second C3b is then added to make the alternative pathway C5 convertase. Recent investigations have revealed that the alternative pathway can itself be initiated in three distinct ways. The first mode of initiation to be discovered, the “tickover” pathway, utilizes the four serum components C3, factor B, factor D, and properdin (see Table 6-1). The term tickover refers to the fact that C3 is constantly being made and spontaneously inactivated and is thus undergoing “tickover.” Two additional modes of activation for the alternative pathway have also been identified: one is initiated by the protein properdin, and the other by proteases such as thrombin and kallikrein. The story of the discovery of properdin is addressed in Classic Experiment Box 6-1. The Alternative Tickover Pathway The alternative tickover pathway is initiated when C3, which is at high concentrations in serum, undergoes spontaneous hydrolysis at its internal thioester bond, yielding the molecule C3(H2O). The conformation of C3(H2O) has been demonstrated by spectrophotometric means to be different from that of the parent protein, C3. C3(H2O) accounts for approximately 0.5% of plasma C3 and, in the presence of serum Mg2⫹ C3a C3b C3(H2O) Factor B Factor B C3bB C3(H2O)B Factor D Factor D Ba The Alternative Pathway Is Initiated in Three Distinct Ways Fluid phase C3 convertase Ba C3bBb Membrane-bound C3 convertase C3(H2O)Bb Fluid phase C3 convertase (c) C3 convertase Properdin (factor P) + C3 C3b C3a C3bBbC3b C5 convertase C5 C5b C5a FIGURE 6-8 The initiation of the alternative tickover pathway of complement. (a) Spontaneous hydrolysis of soluble C3 to C3(H2O) allows the altered conformation of C3(H2O) to bind factor B, rendering it susceptible to cleavage by factor D. The resulting complex C3(H2O)Bb forms a fluid phase convertase capable of cleaving C3 to C3a and C3b. (b) Some of the C3b molecules formed by the fluid phase convertase bind to cell membranes. C3b, like C3(H2O) binds factor B in such a way as to make B susceptible to factor D-mediated cleavage. (c) However, membrane-bound C3bBb is unstable in the absence of properdin (factor P), which binds the C3bBb complex on the membrane and stabilizes it. Addition of a second C3b molecule to the C3bBb complex forms the C5 convertase, which is also stabilized by factor P. The Complement System the newly formed C3b molecules bind nearby microbial surfaces via their thioester linkages. Since factor B is capable of binding to C3b as well as to C3(H2O), factor B now binds the newly attached C3b molecules on the microbial cell surface (see Figure 6-8b), and again becomes susceptible to cleavage by factor D, with the generation of C3bBb complexes. These C3bBb complexes are now located on the microbial membrane surface. Like the C4b2a complexes of the classical pathway, the cell-bound C3bBb complexes have C3 convertase activity, and this complex now takes over from the fluid phase C3(H2O)Bb as the predominant C3 convertase. To be clear, there are two C3 convertases in the alternative tickover pathway: a fluid phase C3(H2O)Bb, which kicks off the pathway, and a membrane-bound C3bBb C3 convertase. The cell-bound C3bBb C3 convertase is unstable until it is bound by properdin, a protein from the serum (Figure 6-8c). Once stabilized by properdin, these cell-associated, C3bBb C3 convertase complexes rapidly generate large amounts of C3b at the microbial surface; these, in turn, bind more factor B, facilitating its cleavage and activation and resulting in a dramatic amplification of the rate of C3b generation. This amplification pathway is rapid and, once the alternative pathway has been initiated, more than 2  106 molecules of C3b can be deposited on a microbial surface in less than 5 minutes. Just as the C5 convertase of the classical and lectin pathways was formed by the addition of C3b to the C4b2a C3 convertase complex, so the C5 convertase of the alternative pathway is formed by the addition of C3b to the alternative pathway C3 convertase complex. The C5 convertase complex therefore has the composition C3bBbC3b, and, like the C3 convertase, is also stabilized by binding to factor P. Like the classical and lectin pathway C5 convertase, C3bBbC3b cleaves C5, which goes on to form the MAC (see Figure 6-8c). The Alternative Properdin-Activated Pathway In the previous section, we introduced properdin as a regulatory factor that stabilizes the C3bBb, membrane-bound C3 convertase. However, recent data suggest that, in addition to stabilizing the ongoing activity of the alternative pathway, properdin may also serve to initiate it. In vitro experiments demonstrated that if properdin molecules were attached to an artificial surface and allowed to interact with purified complement components in the presence of Mg2, the immobilized properdin bound C3b and factor B (Figure 6-9). This bound factor B proved to be susceptible to cleavage by factor D, and the resultant C3bPBb complex acted as an effective C3 convertase, leading to the amplification process discussed above. Thus, it seemed that properdin could initiate activation of the alternative pathway on an artificial substrate. However, proving that a set of reactions can occur in vitro does not necessarily mean that it actually does occur in vivo. Investigators next proved properdin’s ability to bind specifically to certain microbes, including Neisseria gonorrhoeae, as Properdin C3b | CHAPTER 6 Factor B 197 Factor D C3bPB C3bPBb Ba FIGURE 6-9 Initiation of the alternative pathway by specific, noncovalent binding of properdin to the target membrane. Properdin binds to components of microbial membranes, and stabilizes the binding of C3bBb complexes of the alternative complement pathway. The difference between this and the tickover pathways is that properdin binds first and initiates complement deposition on the membrane. well as to apoptotic and necrotic cell surfaces. Once bound, properdin was indeed able to initiate the alternative pathway, as indicated above. Support for the physiological relevance of this properdininitiated pathway is provided by the observation that patients with a deficiency in properdin production are uniquely susceptible to meningococcal disease induced by the Neisseria gonorrhoeae bacterium. These findings suggest that properdin has the capacity to act as a pattern recognition receptor (PRR), specifically directing the activation of the alternative pathway onto the surface of Neisseria and other microbial cells. Note that this pathway relies on the preexistence of low levels of C3b, which must be generated by mechanisms such as the tickover pathway. However, the specific binding of properdin to Neisseria membranes demonstrates how the properdin pathway can provide greater selectivity than that available from the nonspecific C3b binding of the tickover pathway. The Alternative Protease-Activated Pathway The complement and blood-coagulation pathways both use protease cleavage and conformational alterations to modify enzyme activities, as well as amplification of various steps of the pathways by feed-forward loops. Recently, some elegant work has revealed the existence of functional interactions as well as theoretical parallels between these two proteolytic cascades. Several decades ago, it was shown that protein factors involved in blood clotting, such as thrombin, could cleave the complement components C3 and C5, in vitro, with the release of the active anaphylatoxins C3a and C5a. Since these cleavage reactions required relatively high thrombin concentrations, they were at first thought not to be physiologically meaningful. More recently, however, it has been demonstrated in a mouse disease model that initiation of the coagulation cascade may result in the cleavage of physiologically relevant amounts of C3 and C5 to produce C3a and C5a. Specifically, in an immune-complex model of acute lung inflammation, thrombin was shown to be capable of cleaving C5, with the release of active C5a. This thrombin-mediated C5 cleavage was also demonstrated in C3 knockout mice, in which the canonical C5 convertases could not possibly have been generated. Thus, strong inflammatory reactions can result in the activation of at least a part of the alternative 198 PA R T I I | Innate Immunity CLASSIC EXPERIMENT The Discovery of Properdin The study of the history of science shows us that scientists, like any other professionals, are often tempted to think about problems only in ways that are already well trodden and familiar. Science, like art and fashion, has its fads and its “in crowds.” Indeed, sometimes those whose work moves in directions too different from that of the mainstream in their field have difficulty gaining credibility until the thinking of the rest of their colleagues catches up with their own. Such was the case for Louis Pillemer, the discoverer of properdin; by the time others in the burgeoning field of immunochemistry appreciated the power of his discovery, it was quite simply too late. Louis Pillemer (Figure 1) was born in 1908 in Johannesburg, South Africa. In 1909 the family emigrated to the United States and settled in Kentucky. Pillemer completed his bachelor’s degree at Duke University and started medical school at the same institution. However, in the middle of his third year, the emotional problems that would plague him for the rest of his life surfaced for the first time, and he left. At that time, deep in the Depression, individuals in Kentucky who could pass an exam in the rudiments of medical care were encouraged to care for patients not otherwise served by a physician. Pillemer dutifully passed this examination and began to travel through Kentucky on horseback, visiting the sick and tendering whatever treatments were then available. In 1935, he quit this wandering life and entered graduate school at what was then Western Reserve University (now Case Western Reserve University). There, Pillemer earned his Ph.D. and, except for some time at Harvard and a tour of duty at the Army Medical School in Washington D.C., he remained at Case Western Reserve for the rest of his life, developing a reputation as an excellent biochemist. Among his more noteworthy accomplishments were the first purifications of both tetanus and diphtheria toxins, which were used, along with killed pertussis organisms, in the development of the standard DPT vaccine. FIGURE 1 Louis Pillemer, the discoverer of properdin. [Presidential Address to the AAI, 1980, by Lepow; reprinted in J. Immunol. 125, 471, 1980. Fig.1. 1980. The American Association of Immunologists, Inc.] After these successes, Pillemer turned his attention to the biochemistry of the complement system, which he had initially encountered during his graduate work. The antibody-mediated classical pathway had already been established. However, Pillemer was intrigued by some more recent experiments that had shown that mixing human serum with zymosan, an insoluble carbohydrate extract from yeast cell walls, at 37C resulted in the selective loss of the vital third component of complement, C3. He was curious about the mechanism of this loss of C3, and initially interpreted the result to suggest that the C3 was being selectively adsorbed onto the zymosan surface. He reasoned that, if this were true, adsorption to zymosan might be used as a method to purify C3 from plasma. However, that first idea was proven incorrect, and instead he began to investigate whether the loss of C3 resulted from C3 cleavage that was occurring at the zymosan surface. Pillemer succeeded in demonstrating that such was indeed the case and went on to show that C3 cleavage only happened when his experiments were run at pH 7.0 and at 37C. This suggested to him that perhaps an enzyme in the serum was binding to the zymosan and causing the inactivation of the C3 component. Consistent with this hypothesis, when he mixed the serum and the zymosan at 17C, no cleavage occurred. However, if he allowed the serum and zymosan to mix at 17C and then warmed up the mixture to 37C, the C3 was cleaved as effectively as if it had been incubated at 37C all along. Next, he incubated the serum and zymosan together at 17C, spun down and removed the zymosan from the mixture, and then added fresh zymosan to the remaining serum containing the C3. He then raised the temperature to 37C. Nothing happened. The C3 was untouched. Whatever enzymic activity in the serum was responsible for the breakdown of C3 The Complement System | CHAPTER 6 199 BOX 6-1 had been adsorbed by the zymosan and removed from the serum. Pillemer concluded that a factor present in serum and adsorbed onto the zymosan was necessary for the cleavage of C3. With his students and collaborators, he purified this component and named it properdin, from the Latin perdere meaning “to destroy.” His flow sheet for these initial experiments, and published as part of his landmark Science paper of 1954, is shown in Figure 2. He also identified a heat-labile factor in the serum that was required for C3 cleavage to occur. Pillemer and colleagues went on to characterize properdin as a protein that represented less than 0.03% of serum proteins and whose activity was absolutely dependent on the presence of magnesium ions. In a brilliant series of experiments, Pillemer demonstrated the importance of his newly discovered factor in complement-related antibacterial and antiviral reactions, as well as its role in the disease known as paroxysmal nocturnal hemoglobinemia. In light of today’s knowledge, we can interpret exactly what was happening in his experiments. Properdin bound to the zymosan and stabilized the C3 convertase of the alternative pathway, resulting in C3 cleavage. Indeed, experiments performed as recently as 2007 show that properdin binds to zymosan in a manner similar to its binding to Neisseria membranes. Pillemer’s discovery of properdin coupled with his purification of the tetanus and diphtheria toxins should have sealed his reputation as a world-class biochemist. Indeed, his findings were deemed of sufficient general interest at the time that they were publicized in the lay press as well as in the scientific media; articles and editorials about them appeared in the New York Times, Time magazine, and Collier’s. Nor did Pillemer oversell his case. Scientists writing about this period are at pains to note that Pillemer did not make or broadcast any claims for his molecule beyond what appeared in the reviewed scientific literature. However, in 1957 and 1958, scientist Robert Nelson offered an alternative explanation for Pillemer’s findings. He pointed out that what Pillemer had described as a new protein could simply be a mixture of natural antibodies specific for zymosan. If that were the case, then all Pillemer had succeeded in doing was describing the classical pathway a FIGURE 2 The flow sheet of Pillemer’s experiments, which showed that a substance in serum that is adsorbed onto zymosan, the yeast cell wall extract, is capable of catalyzing the cleavage of C3. [Pillemer, L., et al. The properdin system and immunity. I. Demonstration and isolation of a new serum protein, properdin, and its role in immune phenomena. Science. 1954 Aug 20;120(3112):279-85. Reprinted with permission from AAAS.] second time. Sensitive biochemical experiments indeed demonstrated the presence of low levels of antizymosan antibodies in properdin preparations, and the immunological community began to doubt the relevance of properdin to the complement activation that Pillemer described. Pillemer was devastated. Never completely stable emotionally, Pillemer’s “behavior became erratic, he occasionally abused alcohol and he appeared to be experimenting with drugs,” according to his former graduate student, Irwin H. Lepow, who went on to become a distinguished immunologist in his own right. On August 31, 1957, right in the midst of the controversy over properdin, Pillemer died of acute barbiturate intoxication. His death was ruled a suicide, although nobody can know whether he was merely seeking short-term relief from stress and his death was therefore accidental, or whether he truly meant to bring an end to his life. Subsequent experiments demonstrated that antibodies to zymosan could be removed from partially purified properdin without loss of the ability of the preparation to catalyze the cleavage of C3, thus confirming Pillemer’s finding. Furthermore, the heat-labile factor identified by Pillemer’s experiments was identified as a previously unknown molecule, which was subsequently named factor B. By the late 1960s other laboratories entered the arena, and slowly Pillemer’s discovery was confirmed and extended into what we now know as the alternative pathway of complement activation. It is one of the great tragedies of immunology that Pillemer did not live to enjoy the validation of his elegant work. Lepow, I. H. 1980. Louis Pillemer, properdin and scientific controversy. Presidential address to American Association of Immunologists in Anaheim, CA, April 16, 1980. Journal of Immunology 125:471–478. Pillemer, L., et al. 1954. The properdin system and immunity. I. Demonstration and isolation of a new serum protein, properdin, and its role in immune phenomena. Science 120:279–285. PA R T I I 200 | Innate Immunity complement pathway, via the enzymes of the coagulation cascade. Given the proinflammatory roles of the anaphylatoxins, this would result in further amplification of the inflammatory state. Additional experiments have since revealed that other coagulation pathway enzymes, such as plasmin, are capable of generating both C3a and C5a. Furthermore, when blood platelets are activated during a clotting reaction, they release high concentrations of ATP and Ca2⫹ along with serine/threonine kinases. These enzymes act to phosphorylate extracellular proteins, including C3b. Phosphorylated C3b is less susceptible to proteolytic degradation than its unphosphorylated form, and thus, by this route, activation of the clotting cascade enhances all of the complement pathways. The Three Complement Pathways Converge at the Formation of the C5 Convertase The end result of the three initiation pathways is the formation of a C5 convertase. For the classical and lectin pathways, the C5 convertase has the composition C4b2a3b; for the alternative pathway, the C5 convertase has the formulation C3bBbC3b. (In the thrombin-initiated pathway, the anaphylatoxin C5a is formed by cleavage of C5 by the blood clotting enzymes, but functionally meaningful C5b concentrations are not generated by this route.) However, the end result of all types of C5 convertase activity is the same: the cleavage of the C5 molecule into two fragments, C5a and C5b. The large C5b fragment is generated on the surface of the target cell or immune complex and provides a binding site for the subsequent components of the MAC. However, the C5b component is extremely labile, is not covalently bound to the membrane like C3b and C4b, and is rapidly inactivated unless it is stabilized by the binding of C6. C5 Initiates the Generation of the MAC Up to this point in the complement cascades, all of the complement reactions take place on the hydrophilic surfaces of microbes or on immune complexes in the fluid phase of blood, lymph, or tissues. In contrast, when C5b binds C6 and C7, the resulting complex undergoes a conformational change that exposes hydrophobic regions on the surface of the C7 component capable of inserting into the interior of the microbial membrane (Figure 6-10a). If, however, the reaction occurs on (a) C7 C9 C9 C8 C9 C5b C6 C5b C5b C6 C7 C5b C6 C7 C8 C5b C6 C7 C8 C9 C6 C7 C8 C9 C9 C9 C9 C9 9 Cell membrane (b) FIGURE 6-10 Formation of the membrane attack complex (MAC). (a) The formation of the MAC showing the addition of C6, C7, C8, and C9 components to the C5b component. (b) Photomicrograph of poly-C9 complex formed by in vitro polymerization of C9 (inset) and complement-induced lesions on the membrane of a red blood cell. These lesions result from formation of membrane-attack complexes. [Part b from E. R. Podack, 1986, Immunobiology of the Complement System, G. Ross, ed., Academic Press, Orlando, FL; part b inset from J. Humphrey and R. Dourmashkin, 1969, The lesions in cell-membranes caused by complement, Advances in Immunology, 11:75.] The Complement System an immune complex or other noncellular activating surface, then the hydrophobic binding sites cannot anchor the complex and it is released. Released C5b67 complexes can insert into the membrane of nearby cells and mediate “innocent bystander” lysis. Under physiologic conditions, such lysis is usually minimized by regulatory proteins (see below). C8 is made up of two peptide chains: C8 and C8. Binding of C8 to the C5b67 complex induces a conformational change in the C8 dimer such that the hydrophobic domain of C8 can insert into the interior of the phospholipid membrane. The C5b678 complex can create a small pore, 10 Å in diameter, and formation of this pore can lead to lysis of red blood cells, but not of nucleated cells. The final step in the formation of the MAC is the binding and polymerization of C9 to the C5b678 complex. As many as 10 to 19 molecules of C9 can be bound and polymerized by a single C5b678 complex. During polymerization, the C9 molecules undergo a transition, so that they, too, can insert into the membrane. The completed MAC, which has a tubular form and functional pore diameter of 70 Å to 100 Å, consists of a C5b678 complex surrounded by a poly-C9 complex (Figure 6-10b). Loss of plasma membrane integrity leads irreversibly to cell death. The Diverse Functions of Complement Table 6-2 lists the main categories of complement function, each of which is discussed below. In addition to its longunderstood role in antibody-induced lysis of microbes, TABLE 6-2 | CHAPTER 6 201 complement also has important functions in innate immunity, many of which are mediated by the soluble innate immune receptors such as MBL and the ficolins. The pivotal importance of C3b-mediated responses such as opsonization has been clearly demonstrated in C3/ knockout animals, which display increased susceptibility to both viral and bacterial infections. In addition, recent experiments have explored the roles of various complement components at the interface of innate and adaptive immunity, and identified multiple mechanisms by which the release of active complement fragments acts to regulate the adaptive immune system. Complement also plays an important role in the contraction phase of the adaptive immune response, and recent work has even suggested that it is important in the elimination of excess synapses during the development of the nervous system. These various functions are described below. Complement Receptors Connect ComplementTagged Pathogens to Effector Cells Many of the biological activities of the complement system depend on the binding of complement fragments to cell surface complement receptors. In addition, some complement receptors play an important role in regulating complement activity by mediating proteolysis of biologically active complement components. The levels of a number of the complement receptors are subject to regulation by aspects of the innate and adaptive immune systems. For example, activation of phagocytic cells by various agents, including the The three main classes of complement activity in the service of host defense Activity Responsible complement component Innate defense against infection Lysis of bacterial and cell membranes Opsonization Induction of inflammation and chemotaxis by anaphylatoxins Membrane attack complex (C5b–C9) Covalently bound C3b, C4b C3a, C4a, C5a (anaphylatoxins) and their receptors on leukocytes Interface between innate and adaptive immunity Augmentation of antibody responses Enhancement of immunologic memory Enhancement of antigen presentation Potential effects on T cells C3b and C4b and their proteolyzed fragments bound to immune complexes and antigen; C3 receptors on immune cells C3b and C4b and their fragments bound to antigen and immune complexes; receptors for complement components on follicular dendritic cells MBL, C1q, C3b, C4b, and C5a C3, C3a, C3b, C5a Complement in the contraction phase of the immune response Clearance of immune complexes from tissues Clearance of apoptotic cells Induction of regulatory T cells C1, C2, C4; covalently bound fragments of C3 and C4 C1q; covalently bound fragments of C3 and C4. Loss of CD46 triggers immune clearance. CD46 This table has been considerably modified from the superb formulation of Mark Walport, Complement, New England Journal of Medicine 344:1058–1066, Table 1. 202 TABLE 6-3 PA R T I I | Innate Immunity Receptors that bind complement components and their breakdown products Receptor Alternative name(s) Ligand Cell surface binding or expression Function CR1 CD35 C3b, C4b, C1q, iC3b Erythrocytes, neutrophils, monocytes, macrophages, eosinophils, FDCs, B cells, and some T cells Clearance of immune complexes, enhancement of phagocytosis, regulation of C3 breakdown CR2 CD21, Epstein-Barr virus receptor C3d, C3dg (human), C3d (mouse) iC3b B cells and FDCs Enhancement of B-cell activation, B-cell coreceptor, and retention of C3d-tagged immune complexes CR3 CD11b/CD18, Mac-1 iC3b and factor H Monocytes, macrophages, neutrophils, NK cells, eosinophils, FDCs, T cells Binding to adhesion molecules on leukocytes, facilitates extravasation; iC3b binding enhances opsonization of immune complexes CR4 CD11c/CD18 iC3b Monocytes, macrophages, neutrophils, dendritic cells, NK cells, T cells iC3b-mediated phagocytosis CRIg VSIG4 C3b, iC3b, and C3c Fixed-tissue macrophages iC3b-mediated phagocytosis and inhibition of alternative pathway C1qRp CD93 C1q, MBL Monocytes, neutrophils, endothelial cell, platelets, T cells Induces T-cell activation; enhances phagocytosis SIGN-R1 CD209 C1q Marginal zone and lymph node macrophages Enhances opsonization of bacteria by MZ macrophages C3aR None C3a Mast cells, basophils, granulocytes Induces degranulation C5aR CD88 C5a Mast cells, basophils, granulocytes, monocytes, macrophages, platelets, endothelial cells, T cells Induces degranulation; chemoattraction; acts with IL-1 and/or TNF- to induce acute phase response; induces respiratory burst in neutrophils C5L2 None C5a Mast cells, basophils, immature dendritic cells Uncertain, but most probably downregulates proinflammatory effects of C5a After Zipfel, P. F., and C. Skerka, 2009, Complement regulators and inhibitory proteins. Nature Reviews Immunology 10:729–740; Kemper, C., and J. P. Atkinson, 2007, T-cell regulation: With complements from innate immunity, Nature Reviews Immunology 7:9–18; Eggleton, P., A. J. Tenner, and K. B. M. Reid, 2000, C1q receptors, Clinical and Experimental Immunology 120:406–412; Ohno, M., et al., 2000, A putative chemoattractant receptor, C5L2, is expressed in granulocyte and immature dendritic cells, but not in mature dendritic cells, Molecular Immunology 37(8):407–412; and Kindt, T. J., B. A. Osborne, and R. A. Goldsby, 2006, Kuby Immunology, 6th ed., New York: W. H. Freeman, Table 7-4. anaphylatoxins of the complement system, has been shown to increase the number of complement receptors by as much as tenfold. Therefore, before we venture into a discussion of the biological functions of complement, we should first become familiar with the receptors for complement components and their activities. The complement receptors and their primary ligands are listed in Table 6-3. Where receptors have more than one name, we have offered both on first introduction, and subsequently refer to that receptor by the more common name. CR1 (CD35) is expressed on both leukocytes and erythrocytes and binds with high affinity to C4b, C3b, and smaller C3b breakdown products. CR1 receptors on erythrocytes bind immune complexes and take them to the liver where they are picked up by phagocytes and cleared. Binding of complement-opsonized microbial cells via CR1 on phagocytes results in receptor-mediated phagocytosis and the secretion of proinflammatory molecules such as IL-1 and prostaglandins. CR1 on B cells mediates antigen uptake of C3b-bound antigen, leading to its degradation in the B-cell lysosomal system and subsequent presentation to T cells. This process makes CR1 an actor in the adaptive, as well as in the innate immune response. As we will see later, CR1 also mediates the protection of host cells against the ravages of complement attack, by serving as a cofactor for the destructive cleavage of C3b and C4b on host cell membranes by factor I, as well as acting as an accelerator of the decay of the C3 and C5 convertases. C3b, either in solution or bound to the surface of cells, is subject to breakdown by endogenous proteases. CD21 (CR2) is expressed on B cells and binds specifically to the The Complement System | CHAPTER 6 203 C3d Antigen CD21 (CR2) mIgM CD19 CD81 (TAPA-1) FIGURE 6-11 Coligation of antigen to B cells via IgM and CD21. The B-cell coreceptor is a complex of the three cell membrane molecules, CD21 (CR2), CD81 (TAPA-1), and CD19. The CD19 component is important in BCR antigen receptor signaling. Antigen that has been covalently bound to fragments of the C3 complement compo- breakdown products of C3b: iC3b, C3d, and C3dg. Since C3b can form covalent bonds with antigens, the presence of CD21 on B cells enables the B cell to bind antigen via both the B-cell receptor and CD21 (Figure 6-11). This ability to simultaneously coengage antigen through two receptors has the effect of reducing the antigen concentration necessary for B-cell activation by up to a hundredfold (see Chapter 12). Deficiencies in CD21 have been identified in patients suffering from autoimmune diseases such as systemic lupus erythematosus (see below). CR3 (a complex of CD11b and CD18) and CR4 (a complex of CD11c and CD18) are important in the phagocytosis of complement-coated antigens. CR3 and CR4 bind C3b and several of its breakdown products, including iC3b, C3c, and C3dg. CRIg also binds C3b. It is expressed on macrophages resident in the fixed tissues, including on the Kupffer cells of the liver. Its importance in clearing C3b-opsonized antigens by facilitating their removal from circulation in the liver is emphasized by the finding that CRIg-deficient mice are unable to efficiently clear C3-opsonized particles. Animals with this deficiency are therefore subject to higher mortalities during infections. C3aR, C5aR, and C5L2 are all members of the G protein coupled receptor (GPCR) family first encountered in Chap- nent is bound by both the immunoglobulin BCR and by the CD21 complement receptor, thus significantly increasing the avidity of the cell receptors for the antigen and allowing lower concentrations of antigen to trigger B-cell activation. ter 4. C3aR and C5aR mediate inflammatory functions after binding the small anaphylatoxins, C3a and C5a, respectively (Figure 6-12). The C5L2 receptor also binds C5a, is structurally similar to C5aR, and is expressed on some of the same cells. However, C5L2 is not functionally coupled to the G protein signaling pathway used by C5aR; instead, signaling through C5L2 appears to modulate C5a signaling through C5aR and C5L2 knockout animals express enhanced inflammatory responses upon C5a binding to its receptor. More recently, a transmembrane lectin, SIGN-R1, able to bind C1q has been shown to be expressed on the surface of macrophages located in the marginal zone of the spleen. SIGN-R1 exists on the macrophage cell surface in aggregated form and is also able to bind carbohydrates present on the coat of the Gram-positive bacterium Staphylococcus pneumoniae. When SIGN-R1 binds to bacterial polysaccharides, the C1q-binding capacity is activated in the same, or in nearby SIGN-R1 molecules, eventually resulting in the opsonization of the bacterium with C3b. The opsonized bacteria are then released from these macrophages and bound by nearby phagocytes, B cells, or dendritic cells. This unusual mechanism explains a problem long noted with patients who have undergone splenectomy: an increased susceptibility to infection by S. pneumoniae. PA R T I I 204 | Innate Immunity Anaphylatoxins and inflammatory response C3a Chemotaxis Cytokine production Phagocytosis C5a Macrophage NH2 NH2 or C3aR C5aR Neutrophil Chemotaxis Oxidative burst Phagocytosis Degranulation Degranulation Basophil COOH COOH Degranulation Chemotaxis α β α γ GDP GTP GDP Eosinophil GTP β Degranulation Chemotaxis γ Mast cell FIGURE 6-12 Binding of the anaphylatoxins C3a and C5a to the G-protein-coupled receptors C3aR and C5aR. The C3aR and C5aR receptors are members of the G-protein-coupled receptor family described in Chapter 4. Binding of the anaphylatoxins to these receptors stimulates the release of proinflammatory mediators from macrophages, neutrophils, basophils, eosinophils, and mast cells, as indicated. [Adapted from J. R. Dunkelberger and W.-C. Song, 2010, Complement and its role in innate and adaptive immune responses, Cell Research 20:34–50, Figure 3B.] Complement Enhances Host Defense Against Infection Complement proteins actively engage in host defense against infection by forming the MAC, by opsonizing potentially pathogenic microbes, and by inducing an inflammatory response that helps to guide leukocytes to the site of infection. MAC-Induced Cell Death The first function of complement to be described was its role in inducing cell death following insertion of the MAC into target cell membranes. Early experiments on MAC formation used erythrocytes as the target membranes, and in this cellular system large pores involving 17 to 19 molecules of C9 were reported. Formation of these holes in the cell membrane (see Figure 6-10b) facilitated the free movement of small molecules and ions. The penetrated red blood cell membranes were unable to maintain osmotic integrity, and the cells lysed after massive influx of water from the extracellular fluid. However, subsequent studies using nucleated eukaryotic cells indicated that smaller pores can be generated using only a few molecules of C9, and that death under these circumstances occurs via a type of apoptosis (programmed cell death), following calcium influx into the cytoplasm. Nuclear fragmentation, a hallmark of apoptotic death, was observed during MAC-induced lysis of nucleated cells, which further supported the notion that at least some MAC-targeted cells succumb to apoptosis. More careful observations indicated that the apoptosis induced by the MAC does not share all the biochemical characteristics normally associated with pro- grammed cell death, and so this MAC-induced apoptosis has been termed apoptotic necrosis. Killing eukaryotic cells with complement is actually quite difficult, as the membranes of such cells have a number of factors that act together to inactivate the complement proteins, and thus protect the host cells from collateral damage during a complement-mediated attack on infectious microorganisms (see below). However, when high concentrations of complement components are present, MACs can overwhelm the host defenses against MAC attack, and the resulting cell fragments, if present in sufficiently high concentrations, can induce autoimmunity. Indeed, complement-mediated damage is a problem in several autoimmune diseases, and the complement system is considered a target for therapeutic intervention in autoimmune syndromes. Can a eukaryotic cell recover from a MAC attack? Welldocumented studies have demonstrated that MACs can be removed from the cell surface, either by shedding MACcontaining membrane vesicles into the extracellular fluid, or by internalizing and degrading the MAC-containing vesicles in intracellular lysosomes. If the MAC is shed or internalized soon enough after its initial expression on the membrane, the cell can repair any membrane damage and restore its osmotic stability. An unfortunate corollary of this capacity to recover from MAC attack is that complement-mediated lysis directed by tumor-specific antibodies may be rendered ineffective by endocytosis or shedding of the MAC (see Chapter 19). Different types of microorganisms are susceptible to complement-induced lysis to varying degrees. Antibody and complement play an important role in host defense against viruses and can be crucial both in containing viral spread The Complement System during acute infection and in protecting against reinfection. Most enveloped viruses, including herpesviruses, orthomyxoviruses such as those causing measles and mumps, paramyxoviruses such as influenza, and retroviruses are susceptible to complement-mediated lysis. However, some bacteria are not susceptible to MAC attack. Gram-positive bacteria efficiently repel complement assault, as the complement proteins cannot penetrate the bacterial cell wall to gain access to the membrane beyond. In contrast, when attacking Gram-negative bacteria, complement first permeabilizes the outer layer and, following destruction of the thin cell wall, lyses the inner bacterial membrane. In some cases, the MAC has been found to localize at regions of apposition of the inner and outer cell walls, and breaches them both simultaneously. Neisseria (a) Opsonization and phagocytosis | CHAPTER 6 205 meningitidis is a Gram-negative bacterium that is susceptible to MAC-induced lysis, and patients who are deficient in any of the complement components of the MAC are particularly vulnerable to potentially fatal meningitis caused by this bacterial species. Promotion of Opsonization As described in Chapter 5, the term opsonization refers to the capacity of antibodies and complement components (as well as other proteins) to coat dangerous antigens that can then be recognized by Fc receptors (for antibodies) or complement receptors (for complement components) on phagocytic cells. Binding of complement-coated antigen by phagocytic cells results in complement receptor-mediated phagocytosis and antigen destruction (Figure 6-13). In (b) Coated virus particle C1q MBLmannose Antibody: C3b/ antigen C4b C3b cleavage products (iC3b, C3c, C3dg) FIGURE 6-13 Opsonization of microbial cells by comple- CR1 (CD35) CR3 (CD11b/ CD18) CR4 (CD11c/ CD18) CRlg Fc receptors SIGNR1 ment components and antibodies. (a) Phagocytosis is mediated by many different complement receptors on the surface of macrophages and neutrophils, including CR1, CR3, CR4, and CrIg. Phagocytes, using their Fc receptors, also bind to antigens opsonized by antibody binding. (b) Electron micrograph of Epstein-Barr virus coated with antibody and C3b and bound to the Fc and C3b receptor (CR1) on a B lymphocyte. [Part b from N. R. Cooper and G. R. Nemerow, 1986, Immunobiology of the Complement System, G. Ross, ed., Orlando, FL: Academic Press.] 206 PA R T I I | Innate Immunity addition, complement receptors on erythrocytes also serve to bind immune complexes, which are then transported to the liver for phagocytosis by macrophages (Figure 6-14). Although less visually compelling than MAC formation, opsonization may be the most physiologically important function assumed by complement components. Opsonization with antibody and complement provides critical protection against viral infection. Antibody and complement can create a thick protein coat around a virus that neutralizes viral infectivity by preventing the virus from binding to receptors on the host cell. It also promotes phagocytosis via the complement receptors, followed by intracellular destruction of the digested particle. Promotion of Inflammation So far, we have focused on the roles of the larger products of complement factor fragmentation: C3b and C4b in opsonization and C5b in the formation of the MAC. However, the smaller fragments of C3, C4, and C5 cleavage—C3a, C4a, and C5a—are no less potent and mediate critically important events in immune responses, acting as anaphylatoxins or inducers of inflammation. We will focus here specifically on the activities of C3a and C5a. C3a and C5a are structurally similar, small proteins (about 9 kDa, or 74–77 amino acid residues in size) that promote inflammation and also serve as chemoattractants for certain classes of leukocytes. C3a and C5a bind G-proteincoupled, activating receptors (C3aR for C3a, C5aR for C5a) on granulocytes, monocytes, macrophages, mast cells, endothelial cells, and some dendritic cells (see Figure 6-12). Binding of these anaphylatoxins to their receptors on some cells triggers a signaling cascade that leads to the secretion of soluble mediators such as IL-6 and TNF-. These cytokines in turn induce localized increases in vascular permeability that enable leukocyte migration into the site of infection, and a concomitant increase in smooth muscle motility that helps to propel the released fluid to the site of damage. In addition, these proinflammatory mediators promote phagocytosis of offending pathogens and localized degranulation of granulocytes (neutrophils, basophils, and eosinophils) with the resultant release of a second round of inflammatory mediators, including histamines and prostaglandins. Inflammatory mediators expedite the movement of lymphocytes into neighboring lymph nodes where they are activated by the pathogen. This localized inflammatory response is further supported by systemic effects, such as fever, that decrease microbial viability. The central role of the anaphylatoxins in the promotion of physiologically important inflammatory responses has brought them to the attention of clinicians seeking ways to tamp down the pathological levels of inflammation experienced by patients suffering from diseases such as rheumatoid arthritis and systemic lupus erythematosus (SLE). Clinical Focus Box 6-2 describes some approaches that investigators are taking to explore the complement system as a potential therapeutic target. BLOOD Ig Ag Soluble immune complex Complement activation C3b CR1 Erythrocyte LIVER AND SPLEEN Phagocyte FIGURE 6-14 Clearance of circulating immune complexes by reaction with receptors for complement products on erythrocytes and removal of these complexes by receptors on macrophages in the liver and spleen. Because erythrocytes have fewer receptors than macrophages, the latter can strip the complexes from the erythrocytes as they pass through the liver or spleen. Deficiency in this process can lead to renal damage due to accumulation of immune complexes. The Complement System Complement Mediates the Interface Between Innate and Adaptive Immunities Experiments in the past two or three decades have revealed multiple mechanisms by which components of the complement system modulate adaptive immunity. Many of these findings are extremely recent, and the study of how the binding of complement components and regulatory proteins may affect antigen-presenting cells, T cells, and B cells is still in its early phases. Complement and Antigen-Presenting Cells Dendritic cells (DCs), follicular dendritic cells (FDCs), and macrophages express all of the known complement receptors. When bound to antigens, MBL, C1q, C3b, and C4b are each capable of engaging their respective receptors on antigenpresenting cells during the process of antigen recognition, and signaling through their respective receptors acts to enhance antigen uptake. In addition, signaling of antigen-presenting cells through the C5aR anaphylatoxin receptor has been shown to modulate their migration and affect interleukin production, particularly that of the cytokine IL-12. Production of IL-12 by an antigen-presenting cell normally skews the T-cell response toward the TH1 phenotype (see Chapter 11). Since both induction and suppression of IL-12 production have been detected after activation of the anaphylatoxin receptors, depending upon the route of antigen delivery, the nature of the antigen, and the maturation status of the antigen-presenting cell, we must infer that many signaling pathways are being integrated to arrive at the eventual biological response to antigen and complement components. Complement and B-Cell-Mediated Humoral Immunity In 1972, Pepys showed that depleting mice of C3 impaired their T-dependent antigen-specific B-cell responses, thus implying that complement may participate in the initiation of the B-cell response. It now appears that this seminal observation was the first description of the complement receptor CD21 acting as a coreceptor in antigen recognition. Recall that the C3b molecule is capable of covalently “tagging” antigen by interaction between its reactive thioester group and hydroxyl or amino groups on the surface of the antigen. As described in Chapters 3 and 12, the immunoglobulin receptor is expressed on the B-lymphocyte membrane along with the complement receptor CD21, the signaling protein CD19, and the transmembrane CD81 (tetraspanin, or TAPA-1) molecule (see Figure 6-11). CD21 is a receptor with specificity for both C3b and the products of C3b proteolytic breakdown, C3d (human) or C3dg (mouse). If C3b was originally bound to a microbial surface, the C3d or C3dg fragments remain bound to that surface after C3b proteolysis. A B cell bearing an immunoglobulin receptor specific for that microbe will therefore be able to bind the microbe with both its immunoglobulin receptor and with | CHAPTER 6 207 the associated CD21 complement receptor, thus enhancing the interaction between the B cell and antigen and reducing the amount of antigen that is required to signal a B-cell response by a factor of tenfold to a thousandfold. Complement and T-Cell-Mediated Immunity The mechanisms by which complement affects T-cell responses are not as well characterized as those for B cells and antigenpresenting cells. However, some recent, interesting results have been generated by the study of mice either genetically deficient in one or more complement components or following treatment with complement inhibitors. For example, mice lacking the C3 gene have reduced CD4⫹ and CD8⫹ T cell responses, implying that signaling through C3 enhances T-cell function. Furthermore, mice treated with inhibitors of C5aR signaling produced fewer antigen-specific CD8⫹ T cells, following influenza infection, than wild-type mice, indicating that C5a may act as a costimulator during CD8⫹ T-cell activation, possibly by increasing IL-12 production by antigen-presenting cells, as described above. Both these experiments demonstrate that signaling through complement components may have positive effects on T-cell-mediated adaptive immune responses. The study of T-cell regulation by complement components is in its infancy, but these provocative findings suggest that there is much to learn and that complement may have more powerful effects on adaptive immunity than previously thought. As these effects are further explored, the question of whether complement is acting directly on the T cells, or indirectly, via effects on antigen-presenting cells, is an issue that will need to be clearly defined. Complement Aids in the Contraction Phase of the Immune Response At the close of an adaptive immune response, most of the lymphocytes that were generated in the initial proliferative phase undergo apoptosis (programmed cell death), leaving only a few antigen-specific cells behind to provide immunological memory (see Chapters 11 and 12). In addition, soluble antigen-antibody complexes may still be present in the bloodstream and immune organs. If autoimmune disease is to be avoided, these excess lymphocytes and immune complexes must be disposed of safely, without the induction of further inflammation, and complement components play a major role in these processes. Disposal of Apoptotic Cells and Bodies Apoptotic cells express the phospholipid phosphatidyl serine on the exterior surface of their plasma membranes. In healthy cells, this phospholipid is normally restricted to the cytoplasmic side of the membrane, and this change in its location serves to signal the immune system that the cell is undergoing programmed cell death. Surprisingly, experiments have also demonstrated the presence of nucleic acids on the surface membrane of apoptotic cells. Nuclear fragmentation and 208 PA R T I I | Innate Immunity CLINICAL FOCUS The Complement System as a Therapeutic Target The involvement of complementderived anaphylatoxins in inflammation makes complement an interesting therapeutic target in the treatment of inflammatory diseases, such as arthritis. In addition, autoimmune diseases that potentially result in complement-mediated damage to host cells, such as multiple sclerosis and agerelated macular degeneration, are also potential candidates for complementfocused therapeutic interventions. The last three decades have seen the crystallization and molecular characterization of several complement components, a necessary precursor to the development of designer drugs targeted to specific proteins. At the time of writing, however, just two therapies directed at interfering with complement components are approved for clinical use, although at least one more is currently in clinical trials. Purified C1 esterase inhibitor is currently in use in the clinic as a therapeutic option for the treatment of hereditary angioedema. However, its beneficial effects are most likely the result of the inhibition of excess esterase activation external to the complement system per se. The other successful complementrelated treatment is more specifically directed toward a dysregulation of the complement cascade. Paroxysmal noctur- nal hemoglobinuria (PNH) manifests as increased fragility of erythrocytes, leading to chronic hemolytic anemia, pancytopenia (loss of blood cells of all types), and venous thrombosis (formation of blood clots). The name PNH derives from the presence of hemoglobin in the urine, most commonly observed in the first urine passed after a night’s sleep. The cause of PNH is a general defect in the synthesis of cell-surface proteins, which affects the expression of two regulators of complement: DAF (CD55) and Protectin (CD59). DAF and Protectin are cell-surface proteins that function as inhibitors of complement-mediated cell lysis, acting at different stages of the process. DAF induces dissociation and inactivation of the C3 convertases of the classical, lectin, and alternative pathways (see Figure 6-16). Protectin acts later in the pathway by binding to the C5b678 complex and inhibiting C9 binding, thereby preventing formation of the membrane pores that destroy the cell under attack. Deficiency in these proteins leads to increased sensitivity of host cells to complement-mediated lysis and is associated with a high risk of thrombosis. Protectin and DAF are attached to the cell membrane via glycosylphosphatidyl inositol (GPI) anchors, rather than by DNA cleavage follow quickly after outer membrane phosphatidyl serine expression. Once apoptosis begins, the dying cell is broken down into membrane-bound vesicles termed apoptotic bodies, which express phosphatidyl serine and/or surfacebound DNA on their exterior surfaces. Recent work has demonstrated that C1q binds specifically to DNA. The complement component C1q specifically binds apoptotic bodies and assists in their clearance. When light-sensitive keratinocytes (skin cells) are treated with UVB to initiate apoptosis and then stained with anti-C1q antibody, C1q staining is restricted to the apoptotic blebs, where DNA, exposed on apoptotic membranes, is specifically recognized by C1q (Figure 6-15). With C1q deposition, the classical pathway is initiated and the apoptotic cells are then opsonized by C3b. This, in turn, leads to clearance of the apoptotic cells by phagocytes. stretches of hydrophobic amino acids, as is the case for many membrane proteins. Patients with PNH lack an enzyme, Phosphatidyl Inositol Glycan Class A (PIGA), that attaches the GPI anchors to the appropriate proteins. Patients thus lack the expression of DAF and Protectin needed to protect erythrocytes against lysis. The term paroxysmal refers to the fact that episodes of erythrocyte lysis are often triggered by stress or infection, which result in increased deposition of C3b on host cells. The gene encoding the PIGA enzyme is X-linked. The defect identified in PNH occurs early in the enzymatic pathway leading to formation of a GPI anchor and resides in the pig-a gene. Transfection of cells from PNH patients with an intact pig-a gene restores the resistance of the cells to host complement lysis. Further genetic analysis revealed that the defect is not encoded in the germline genome (and therefore is not transmissible to offspring), but rather is the result of mutations that occurred within the hematopoietic stem cells themselves, such that any one individual may have both normal and affected cells. Those patients who are most adversely affected display a preferential expansion of the affected cell populations. PNH is a chronic disease with a mean survival time between 10 and 15 years. In the absence of C1q, the apoptotic membrane blebs are released from the dying cells as apoptotic bodies, which can then act as antigens and initiate autoimmune responses. As a consequence, mice deficient in C1q show increased mortality and higher titers of auto-antibodies than control mice and also display an increased frequency of glomerulonephritis, an autoimmune kidney disease. Analysis of the kidneys of C1q knockout mice reveal deposition of immune complexes as well as significant numbers of apoptotic bodies. Disposal of Immune Complexes As mentioned earlier, the coating of soluble immune complexes with C3b facilitates their binding by CR1 on erythrocytes. Although red blood cells express lower levels of CR1 The Complement System | CHAPTER 6 209 BOX 6-2 The most common causes of mortality in PNH are venous thromboses affecting hepatic veins and progressive bone marrow failure. A breakthrough in treatment of PNH was reported in 2004 using a humanized monoclonal antibody that targets complement component C5 and thus inhibits the terminal steps of the complement cascade and formation of the membraneattack complex. This antibody—eculizumab (Soliris)—was infused into patients, who were then monitored for the loss of red blood cells. A dramatic improvement was seen in patients during a 12-week period of treatment with eculizumab (Figure 1). Treatment of PNH patients with eculizumab relieves hemoglobinuria, reverses the kidney damage resulting from high levels of protein in the urine, and significantly reduces the frequency of thromboses. In 2007, eculizumab was approved for the treatment of PNH in the general population. Since the control of infections with Meningococcal bacteria (Neisseria meningitidis) relies upon an intact and functioning membrane attack complex (MAC), patients being treated with eculizumab are routinely vaccinated against meningococcus. The pathology of PNH underscores the potential danger to the host that is inherent in the activation of the complement Days with paroxysms per patient per month 3.0 2.5 2.0 1.5 1.0 0.5 0 Before administration of eculizumab During administration of eculizumab FIGURE 1 Treatment of PNH patients with eculizumab relieves hemoglobinuria. The number of days with paroxysm (onset of attack) per patient per month in the month prior to treatment (left bar) and for a 12-week period of treatment with eculizumab (right bar) is shown. [From P. Hillmen, et al., 2004. Eculizumab in patients with paroxysmal nocturnal hemoglobinuria. New England Journal of Medicine 350:6, 552–559.] (100–1000 molecules per cell, depending on the age of the cell and the genetic constitution of the donor) than do granulocytes (5  104 per cell), there are about 1000 erythrocytes for every white blood cell, and therefore erythrocytes account for about 90% of the CR1 in blood. Erythrocytes also therefore play an important role in clearing C3b-coated immune complexes by conveying them to the liver and spleen where the immune complexes are stripped from the red blood cells and phagocytosed (see Figure 6-14). The importance of the complement system in immune complex clearance is emphasized by the finding that patients with SLE, the autoimmune disease, have high concentrations of immune complexes in their serum that are deposited in system. Complex systems of regulation are necessary to protect host cells from the activated complement complexes generated to lyse intruders, and alterations in the expression or effectiveness of these regulators has the potential to result in a pathological outcome. The centrality of C3 within the complement cascades may suggest that it would be an excellent target for therapeutic intervention in complementmediated disease. However, its very position at the crossroads of the three pathways means that interference with C3 activity places patients at increased risk for infectious diseases normally curtailed by the activities of the innate as well as the adaptive immune systems, and so systemic drugs specifically targeting C3 may prove to have too high a risk-benefit ratio. However, in one success story, targeted to a particular organ system, the drug Compstatin was developed as a result of phage display experiments designed to discover compounds that bound to C3. Chemical refinements based on structural and computational studies followed, and a Compstatin derivative, POT-4, is now in clinical trials for the treatment of age-related macular degeneration, a progressive and debilitating eye disease that leads inexorably and quickly to total blindness. Because POT-4 can be delivered directly into the vitreous humor of the eye, systemic side effects on the patient’s immune system are minimized. the tissues. Complement is activated by these tissue-deposited immune complexes, and pathological inflammation is induced in the affected tissues (see Chapter 15). Since complement activation is implicated in the pathogenesis of SLE, it may therefore seem paradoxical that the incidence of SLE is highly correlated to C4 deficiency. Indeed, 90% of individuals who completely lack C4 develop SLE. The resolution to this paradox lies in the fact that deficiencies in the early components of complement lead to a reduction in the levels of C3b that are deposited on the immune complexes. This reduction, in turn, inhibits their clearance by C3b-mediated opsonization and allows for the activation of the later inflammatory and cytolytic phases of complement activation. 210 PA R T I I | Innate Immunity levels of complement components in the adult retina during the early phases of the disease, suggesting that inappropriate, complement-mediated synaptic pruning may be a causal factor in that disease. C1q up-regulation has also been observed in animal models of both ALS and Alzheimer’s disease, potentially implicating inappropriate complement activation in the etiology of some clinically important neurodegenerative illnesses. The Regulation of Complement Activity FIGURE 6-15 Binding of C1q to apoptotic keratinocytes (skin cells) is limited to surface apoptotic blebs. Apoptosis was induced in human skin cells by exposure to ultraviolet light. The cells were then incubated in human serum and stained with an antibody to C1q. This is an image derived from confocal microscopy that shows the C1q staining in green, clearly localizing to the apoptotic blebs. The cell itself is visualized by phase contrast microscopy, and the image was generated by merging the phase contrast and immunofluorescent frames. [Korb and Ahearn. 1997. J. Immunol. 158: 4525–4528.] Complement Mediates CNS Synapse Elimination Complement has also been shown to play an important role outside the boundaries of the immune system. In the developing nervous system, growing neurons first form a relatively high number of connections (synapses) with one another; the number of those synapses is then pruned as the nervous system matures. Scientists studying the development of the mouse eye have demonstrated that animals deficient in C1q or C3 fail to eliminate these early, excess synapses, and display anatomical abnormalities in the visual nervous system, indicating that complement may play an important role in the process of synaptic remodeling. In healthy animals, the expression of complement components in mouse neurons is up-regulated early in development, in response to signals provided by immature astrocytes (glial cells that assist in the function and maintenance of the nervous system). Complement component expression is then down-regulated as the animals mature and astrocyte function is dampened. However, patients suffering from glaucoma have been shown to display erroneously high All biological systems with the potential to damage the host, be they metabolic pathways, cytotoxic cells, or enzyme cascades, are subject to rigorous regulatory mechanisms, and the complement system is no exception to this general rule. Especially in light of the potent positive feedback mechanisms and the absence of antigen specificity of the alternative pathway, the host must ensure that the destructive potential of complement proteins is confined to the appropriate pathogen surfaces and that collateral damage to healthy host tissue is minimized. Below, we discuss the different mechanisms by which the host protects itself against the potential ravages of inadvertent complement activation. Complement Activity Is Passively Regulated by Protein Stability and Cell-Surface Composition Protection of vertebrate host cells against complementmediated damage is achieved by both general, passive and specific, active regulatory mechanisms. The relative instability of many complement components is the first means by which the host protects itself against inadvertent complement activation. For example, the C3 convertase of the alternative pathway, C3bBbC3b, has a half-life of only 5 minutes before it breaks down, unless it is stabilized by reaction with properdin. A second passive regulatory mechanism depends upon the difference in the cell surface carbohydrate composition of host versus microbial cells. For example, fluid phase proteases that destroy C3b bind much more effectively to host cells, which bear high levels of sialic acid, than to microbes that have significantly lower levels of this sugar. Hence, any C3b that happens to alight on a host cell is likely to be destroyed before it can effect significant damage. In addition to these more passive environmental brakes on inappropriate complement activation, a series of active regulatory proteins act to inhibit, destroy, or tune down the activity of complement proteins and their active fragments. The stages at which complement activity is subject to regulation are illustrated in Figure 6-16, and the regulatory proteins are listed in Table 6-4. The Complement System | CHAPTER 6 211 (a) Dissociation of C1 components C1INH C1r2s2 C1r2s2 C1q (b) Decay-accelerating activity for C3 convertases C4b2a DAF (CD55) CR1 (CD35) C4BP Classical pathway C3bBb DAF (CD55) CR1 (CD35) Factor H Alternative pathway C4b C2a C3b Bb (c) Factor I cofactor activity Factor I C3b MCP (CD46) CR1 (CD35) Factor H Factor I MCP (CD46) CR1 (CD35) C4BP C4b C3c iC3b C4c C4d (d) Inhibition of lysis C5bC6 C7 C8 C3dg C9 Protectin (CD59) Vitronectin/S protein MAC inhibition prevents C9 binding and polymerization (e) Cleavage of the anaphylatoxins C3a des Arg C3a C5a Carboxypeptidase N Carboxypeptidase B Carboxypeptidase R C5a des Arg FIGURE 6-16 The regulation of complement activity. The various stages at which complement activity is subject to regulation (see text for details). [Adapted from J. R. Dunkelberger and W.-C. Song, 2010. Cell Research 20:34–50.] The C1 Inhibitor, C1INH, Promotes Dissociation of C1 Components to control the time period during which they can remain active. C1INH, the C1 inhibitor, is a plasma protein that binds in the active site of serine proteases, effectively poisoning them. C1INH belongs to the class of proteins called serine protease inhibitors (serpins), and it acts by forming a complex with the C1 proteases, C1r2s2, causing them to dissociate from C1q and preventing further activation of C4 or C2 (see Figure 6-16a). C1INH inhibits both C3b and the serine protease MASP2. It is the only plasma protease capable of inhibiting the initiation of both the classical and lectin complement pathways. Its presence in plasma serves Decay Accelerating Factors Promote Decay of C3 Convertases Since the reaction catalyzed by the C3 convertase enzymes is the major amplification step in complement activation, the generation and lifetimes of the C3 convertases C4b2a and C3bBb are subject to rigorous control. Several membrane-bound factors accelerate the decay of the C4b2a on the surface of host cells, including decay accelerating factor, or DAF (CD55), CR1, and C4BP (C4 binding protein) 212 TABLE 6-4 PA R T I I | Innate Immunity Proteins involved in the regulation of complement activity Protein Fluid phase or membrane Pathway affected Function C1 inhibitor (C1INH) Fluid phase Classical and lectin Induces dissociation and inhibition of C1r2s2 from C1q; serine protease inhibitor Decay Accelerating Factor (DAF) CD55 Membrane bound Classical, alternative, and lectin Accelerates dissociation of C4b2a and C3bBb C3 convertases CR1 (CD35) Membrane bound Classical, alternative, and lectin Blocks formation of, or accelerates dissociation of, the C3 convertases C4b2a and C3bBb by binding C4b or C3b Cofactor for factor I in C3b and C4b degradation on host cell surface C4BP Soluble Classical and lectin Blocks formation of, or accelerates dissociation of, C4b2a C3 convertase Cofactor for factor I in C4b degradation Factor H Soluble Alternative Blocks formation of, or accelerates dissociation of, C3bBb C3 convertase Cofactor for factor I in C3b degradation Factor I Soluble Classical, alternative, and lectin Serine protease: cleaves C4b and C3b using cofactors shown in Figure 6-16 Membrane cofactor of proteolysis, MCP (CD46) Membrane bound Classical, alternative, and lectin Cofactor for factor I in degradation of C3b and C4b S protein or Vitronectin Soluble All pathways Binds soluble C5b67 and prevents insertion into host cell membrane Protectin (CD59) Membrane bound All pathways Binds C5b678 on host cells, blocking binding of C9 and the formation of the MAC complex Carboxypeptidases N, B, and R Soluble Anaphylatoxins produced by all pathways Cleave and inactivate the anaphylatoxins C3a and C5a (see Figure 6-16b). These decay accelerating factors cooperate to accelerate the breakdown of the C4b2a complex into its separate components. The enzymatically active C2a diffuses away, and the residual membrane-bound C4b is degraded by another regulatory protein, factor I (see Figure 6-16c). In the alternative pathway, DAF and CR1 function in a similar fashion and are joined by factor H in separating the C3b component of the alternative pathway C3 convertase from its partner, Bb. Again, Bb diffuses away, and residual C3b is degraded (see Figure 6-16b). Whereas DAF and CR1 are membrane-bound components and their expression is therefore restricted to host cells, factor H and C4BP are soluble regulatory complement components. Host-specific function of factor H is ensured by its binding to polyanions such as sialic acid and heparin, essential components of eukaryotic, but not prokaryotic, cell surfaces. Similarly, C4BP is preferentially bound by host cell membrane proteoglycans such as heparan sulfate. In this way, host cells are protected from deposition of complement components, whereas it is permitted on the membranes of microbial invaders that lack DAF and CR1 expression and fail to bind factor H or C4BP. Factor I Degrades C3b and C4b Factor H, C4BP, and CR1 also figure in a second and potentially more critical pathway of complement regulation: that catalyzed by factor I. Factor I is a soluble, constitutively (always) active serine protease that can cleave membraneassociated C3b and C4b into inactive fragments (Figure 6-16c). However, if factor I is indeed soluble, constitutively active, and designed to destroy C3b and C4b, one might wonder how the complement cascades described above can ever succeed in destroying invading microbes? The answer is that factor I requires the presence of cofactors in order to function and that two of these cofactors, Membrane Cofactor of Proteolysis (MCP, or CD46) and CR1, are found on the The Complement System surface of host cells, but not on the surfaces of the microbial cells (see Figure 6-16c). The other two cofactors, factor H and C4BP, described above, bind host cell-surface peptidoglycans. Hence, cleavage of membrane-bound C3b on host cells is conducted by factor I in collaboration with the membrane-bound host cell proteins, CD46 and CR1, and the soluble cofactor H. Similarly, cleavage of membranebound C4b is effected again by factor I, this time in collaboration with membrane-bound CD46 and CR1 and soluble cofactor C4BP. Since these membrane-bound, or membranebinding cofactors are not found on microbial cells, C3b and C4b are thus allowed to remain on microbial cells and effect their specific functions. CD46 has recently been implicated as a factor in the control of apoptosis of dying T cells. When a T cell commits to apoptosis, it expresses DNA on its cell membrane that binds circulating C1q, as described above. It then begins to shed CD46 from the cell surface. Only after CD46 is lost can progression of the classical pathway occur, with increased C3b binding and phagocytosis of the apoptotic T cells. As the cell continues to deteriorate, it may also begin to release nucleotides and other molecules into the tissue fluids that serve as chemotactic signals for phagocytes. Protectin Inhibits the MAC Attack In the case of a particularly robust antibody response, or of an inflammatory response accompanied by extensive complement activation, inappropriate assembly of MAC complexes on healthy host cells can potentially occur, and mechanisms have evolved to prevent the resulting inadvertent host cell destruction. A host cell surface protein, Protectin (CD59), binds any C5b678 complexes that may be deposited on host cells and prevents their insertion into the host cell membrane (see Figure 6-16d). Protectin also blocks further C9 addition to developing MACs. In addition, the soluble complement S protein, otherwise known as Vitronectin, binds any fluid phase C5b67 complexes released from microbial cells, preventing their insertion into host cell membranes. Carboxypeptidases Can Inactivate the Anaphylatoxins C3a and C5a Anaphylatoxin activity is regulated by cleavage of the C-terminal arginine residues from both C3a and C5a by serum carboxypeptidases, resulting in rapid inactivation of the anaphylatoxin activity (Figure 6-16e). Carboxypeptidases are a general class of enzymes that remove amino acids from the carboxyl termini of proteins; the specific enzymes that mediate the control of anaphylatoxin concentrations are carboxypeptidases N, B, and R. These enzymes remove arginine residues from the carboxyl termini of C3a and C5a to form the so-called des-Arg (“without Arginine”), inactive forms of the molecules. In addition, as mentioned above, binding of C5a by C5L2 also serves to modulate the inflammatory activity of C5a. | CHAPTER 6 213 Complement Deficiencies Genetic deficiencies have been described for each of the complement components. Homozygous deficiencies in any of the early components of the classical pathway (C1q, C1r, C1s, C4, and C2) result in similar symptoms, notably a marked increase in immune-complex diseases such as SLE, glomerulonephritis, and vasculitis. The effects of these deficiencies highlight the importance of the early complement reactions in generating C3b and C3b’s role in the solubilization and clearance of immune complexes. In addition, as described above, C1q has been shown to bind apoptotic blebs. In the absence of C1q binding, cells bearing these apoptotic blebs, or the blebs themselves, can act as autoantigens and lead to the development of autoimmune diseases such as SLE. Individuals with deficiencies in the early complement components may also suffer from recurrent infections by pyogenic (pus-forming) bacteria such as streptococci and staphylococci. These organisms are Gram-positive and therefore are normally resistant to the lytic effects of the MAC. Nevertheless, the early complement components can help to mitigate such infections by mediating a localized inflammatory response and opsonizing the bacteria. A deficiency in MBL, the first component of the lectin pathway, has been shown to be relatively common, and results in serious pyrogenic (fever-inducing) infections in babies and children. Children with MBL deficiency suffer from respiratory tract infections. MBL deficiency is also found with a frequency two to three times higher in SLE patients than in normal subjects, and certain mutant forms of MBP are found to be prevalent in chronic carriers of hepatitis B. Deficiencies in factor D and properdin—early components of the alternative pathway—appear to be associated with Neisseria infections but not with immune-complex disease. People with C3 deficiencies display the most severe clinical manifestations of any of the complement deficiency patients, reflecting the central role of C3 in opsonization and in the formation of the MAC. The first person identified with a C3 deficiency was a child suffering from frequent, severe bacterial infections leading to meningitis, bronchitis, and pneumonia, who was initially diagnosed with agammaglobulinemia. After tests revealed normal immunoglobulin levels, a deficiency in C3 was discovered. This case highlighted the critical function of the complement system in converting a humoral antibody response into an effective defense mechanism. The majority of people with C3 deficiency have recurrent bacterial infections and may also present with immune-complex diseases. Levels of C4 (one of the proteins operative very early in the classical pathway) vary considerably in the population. Specifically, the genes encoding C4 are located in the major histocompatibility locus (see Chapter 8), and the number of C4 genes may vary among individuals from two to six. Low gene copy numbers are associated with lower levels of C4 in plasma and with a correspondingly higher incidence of SLE, for the reasons described above. Patients with complete 214 PA R T I I | Innate Immunity deficiencies of the components of the classical pathway such as C4 suffer more frequent infections with bacteria such as S. pneumoniae, Haemophilus influenza, and N. meningitidis. However, even patients with low copy numbers of C4 appear to be relatively well protected against such infections. Interestingly, C4 exists in two isoforms: C4A and C4B. C4B is more effective in binding to the surfaces of the three bacterial species mentioned above, and homozygous C4B deficiency has been shown to result in a slightly higher incidence of infection than is experienced by individuals with C4A deficiency, or those who are heterozygous for the two isoforms. Individuals with deficiencies in members of the terminal complement cascade are more likely than members of the general population to suffer from meningococcal infections, indicating that cytolysis by complement components C5-C9 is of particular relevance to the control of N. meningitidis. This has resulted in the release of public health guidelines that highlight the need for vaccinations against N. meningitidis for individuals deficient in the terminal complement components. Deficiencies of complement regulatory proteins have also been reported. As described above, C1INH, the C1 inhibitor, regulates activation of the classical pathway by preventing excessive C4 and C2 activation by C1. However, as a serine protease inhibitor, it also controls two serine proteases in the blood clotting system. Therefore, patients with C1INH deficiency suffer from a complex disorder that includes excessive production of vasoactive mediators (molecules that control blood vessel diameter and integrity), which in turn leads to tissue swelling and extracellular fluid accumulation. The resultant clinical condition is referred to as hereditary angioedema. It presents clinically as localized tissue edema that often follows trauma, but sometimes occurs with no known TABLE 6-5 cause. The edema can be in subcutaneous tissues; within the bowel, where it causes abdominal pain; or in the upper respiratory tract, where it can result in fatal obstruction of the airway. C1INH deficiency is an autosomal dominant condition with a frequency of 1 in 1000 in the human population. Studies in humans and experimental animals with homozygous deficiencies in complement components have provided important information regarding the roles of individual complement components in immunity. These initial observations have been significantly enhanced by studies using knockout mice, genetically engineered to lack expression of specific complement components. Investigations of in vivo complement activity in these animals has allowed dissection of the complex system of complement proteins and the assignment of precise biologic roles to each. Microbial Complement Evasion Strategies The importance of complement in host defense is clearly illustrated by the number and variety of strategies that have evolved in microbes, enabling them to evade complement attack (see Table 6-5). Gram-positive bacteria have developed thick cell walls and capsules that enable them to shrug off the insertion of MAC complexes, while others escape into intracellular vacuoles to escape immune detection. However, these two general strategies are energy intensive for the microbe, and many microbes have adopted other tactics to escape complement-mediated destruction. In Advances Box 6-3, we describe the multiple ways in which Staphylococcus aureus, a major human pathogen, protects itself against complement-mediated destruction. In this Some microbial complement evasion strategies Complement evasion strategy Example Interference with antibody-complement interaction Antibody depletion by Staphylococcal protein A Removal of IgG by Staphylokinase Binding and inactivation of complement proteins S. aureus protein SCIN binds and inactivates the C3bBb C3 convertase Parasite protein C2 receptor trispanning protein disrupts the binding between C2 and C4 Protease-mediated destruction of complement component Elastase and alkaline phosphatase from Pseudomonas degrade C1q and C3/C3b ScpA and ScpB from Streptococcus degrade C5a Microbial mimicry of complement regulatory components Streptococcus pyogenes M proteins bind C4BP and factor H to the cell surface, accelerating the decay of C3 convertases bound to the bacterial surface Variola and Vaccinia viruses express proteins that act as cofactors for factor I in degrading C3b and C4b The Complement System section, we address complement microbial evasion at a more conceptual level, categorizing the approaches that microbes have evolved to elude this effector arm of the immune response. Some Pathogens Interfere with the First Step of Immunoglobulin-Mediated Complement Activation Many classes of viruses act at the beginning of the classical complement pathway by synthesizing proteins and glycoproteins that specifically bind the Fc regions of antibodies, thus preventing complement binding and the generation of the classical pathway reactions. Some viruses also effect enhanced clearance of antibody-antigen complexes from the surfaces of virus-infected cells and/or manufacture proteins that induce rapid internalization of viral protein-antibody complexes. Microbial Proteins Bind and Inactivate Complement Proteins Since highly specific protein-protein interactions between complement components are central to the functioning of the cascade, it is logical that microbes have evolved mechanisms that interfere with some of these binding reactions. Advances Box 6-3 describes those used by S. aureus, but the strategy of inhibiting the interactions between complement proteins is not restricted to bacteria. The production of molecules that inhibit the interactions between complement components has also been detected in certain human parasites. Specifically, a protein generated by some species of both Schistosoma and Trypanosoma, the complement C2 receptor trispanning protein, disrupts the interaction between C2a and C4b and thus prevents the generation of the classical pathway C3 convertase. Microbial Proteases Destroy Complement Proteins Some microbes produce proteases that destroy complement components. This strategy is utilized primarily by bacteria, and numerous bacterial proteases exist that are capable of digesting a variety of complement components. For example, elastase and alkaline protease proteins from Pseudomonas aeruginosa target the degradation of C1q and C3/C3b, and two proteases derived from Streptococcal bacteria, scpA and scpB, specifically attack the anaphylatoxin C5a. Streptococcal pyrogenic exotoxin B was found to degrade the complement regulator properdin, with resultant destabilization of the alternative pathway C3 convertase on the bacterial surface. Fungi can also inactivate complement proteins. The opportunistic human pathogen Aspergillus fumigatus secretes an alkaline protease, Alp1, that is capable of cleaving C3, C4, C5, and C1q, as well as IgG. Since this pathogen | CHAPTER 6 215 tends to attack patients who are already immunocompromised, its capacity to further damage the immune system is particularly troubling to the clinician. Some Microbes Mimic or Bind Complement Regulatory Proteins Several microbial species have exploited the presence of cellsurface or soluble regulators of complement activity for their own purposes, either by mimicking the effects of these regulators, or by acquiring them directly. Indeed, recent work suggests that sequestration of host-derived regulators of complement activity may be one of the most widely utilized microbial mechanisms for complement evasion. Many microbes have developed the ability to bind one or another of the fluid phase inhibitors of complement, C4BP or factor H. For example, Streptococcus pyogenes, an important human pathogen, expresses a family of proteins called “M proteins” capable of binding to C4BP and factor H. Expression of these regulatory proteins on the bacterial surface results in the inhibition of the later steps of complement fixation. More surprisingly, some microbes appear to acquire membrane-bound regulators from host cells. Helicobacter pylori, obtained from patients suffering from gastric ulcers, was found to stain positive for Protectin, a potent host inhibitor of MAC complex formation. Protectin is normally attached to the host membrane by a glycosylphosphatidyl inositol anchor, and this anchor must be transferable in some way from the host to the bacterial cell membrane. Many viruses have evolved the capacity to produce proteins that closely mimic the structure and function of complement regulatory proteins. For example, the Variola (smallpox) and Vaccinia (cowpox) viruses express complement inhibitory proteins that bind C3b and C4b and serve as cofactors for factor I, thus preventing complement activation at the viral membrane. In addition to generating complement regulatory proteins, it should be noted that some viruses actively induce regulatory complement components within the cells that harbor them. Other viruses camouflage themselves during budding from the host cell by hiding within regions of the host membrane on which regulatory complement components are expressed. And finally, some viruses mimic eukaryotic membranes by incorporating high levels of sialic acid into the viral membrane, thus facilitating the binding of the cofactors for factor I that normally bind only to host cell membranes. This therefore results in the inhibition of complement activation on the viral surface. The Evolutionary Origins of the Complement System Although the complement system was initially characterized by its ability to convert antibody binding into pathogen lysis, studies of complement evolution have identified the PA R T I I 216 | Innate Immunity ADVANCES Staphylococcus aureus Employs Diverse Methods to Evade Destruction by the Complement System Staphylococcus aureus has polysaccharide itself provides the bacterium with some mechanical inhibition of opsonization (1 in Figure 1). Although complement factors can assemble on the cell-wall surface underneath the capsule, most are then inaccessible to the complement receptors on phagocyte sur- developed an impressive variety of mechanisms that inhibit both the classical and the alternative pathways of complement activation. S. aureus is an encapsulated, Grampositive bacterial strain. The capsular faces. However, if the C3 convertases C4b2a and C3bBb do succeed in assembling at the bacterial surface, they are then bound by a small, 9.8 kDa protein called Staphylococcus Complement INhibitor (SCIN). Blockage of further convertase activity prevents amplification of 1 4 Capsule Cell wall Sak Plasminogen 5 ClfA binding fibrinogen Plasmin IgG C3b 2 3 SpA binding IgG Efb C3 C3b C3a components of the lytic MAC as relatively late additions to the animal genome. Long before the emergence of the adaptive immune system, complement components provided invertebrate organisms with an important advance on the peptide-based humoral immunity previously available to them (see Chapter 5). Activation of complement cascades in invertebrates and early vertebrates most likely culminated in opsonization and phagocytosis by primitive hemocytes. It appears probable that, among the non-MAC complement components, the proteins of the alternative pathway were the first to appear, followed by those of the lectin recognition systems. A fully fledged MAC emerged only around the same time as the appearance of the adaptive immune system (see Figure 6-17 and Table 6-6). FIGURE 1 Mechanisms by which S. aureus avoids opsonophagocytosis. (1) The capsular polysaccharides denies access of neutrophils to opsonized bacteria. (2) The extracellular fibrinogen binding protein (Efb) binds C3, preventing it from reaching the cell surface and inhibiting further activation of the complement cascade. (3) Protein A (SpA) binds IgG in a conformation that does not permit Fc receptor binding. (4) Staphylokinase (Sak), secreted by the bacterium, activates plasminogen, a protease capable of cleaving and inactivating IgG and C3b. (5) Clumping factor A binds factor I and localizes it to the microbial surface, where it cleaves and inactivates any C3b that binds there. [Adapted from Foster, T. J., 2005. Immune evasion by Staphylococci. Nature Reviews Microbiology 3: 948–958, Figure 3.] Genomic analysis has classified complement components into five gene families, each of which possesses unique domain structures that have enabled investigators to trace their phylogenetic origins. The first of these gene families encodes C1q, mannose-binding lectin (MBL), and ficolins. Genes for prototypical C1q molecules have been identified in species as primitive as lampreys (the jawless fishlike vertebrate; Phylum Chordata, Subphylum Vertebrata), sea urchins (Phylum Echinodermata, Subphylum Echinozoa), and ascidians (Phylum Chordata, Subphylum Urochordata). Analysis of C1q gene clusters in different species has demonstrated that C1q genes appeared prior to the generation of immunoglobulin genes, and that at least some of these C1q proteins bind to specific carbohydrates, implying that they The Complement System | CHAPTER 6 217 BOX 6-3 the complement cascade. A second protein secreted by S. aureus, the Extracellular Fibrinogen-Binding protein (Efb), binds C3, thus preventing its deposition on the bacterial cell surface (2). But if the immune system succeeds in depositing C3, in spite of all these evasive actions by the bacterium, S. aureus has evolved the capacity to manufacture other proteins to destroy it. Recently, a metalloproteinase enzyme, aureolysin, has been shown to be secreted by S. aureus. Aureolysin cleaves C3 in such a manner that it is then further degraded and inactivated by host regulatory proteins. Other mechanisms interfere with the ability of antibodies to initiate the complement cascade. Staphylococcal protein A (SpA) is anchored into the carbohydrate cell wall, and possesses four or five extracellular domains, each of which is capable of binding to the Fc portion of a molecule of IgG, thus effectively blocking effector functions of the antibody molecules, including complement activation (3). In addition, S. aureus secretes an enzyme called Staphylokinase (SAK) that attaches to the bacterial surface and binds to plasminogen, the host cell plasma zymogen, activating it upon binding (4). The activated plasminogen has a serine protease activity that cleaves and releases any surface-bound IgG and C3b. The S. aureus protein Sbi, which is found both in association with the bacterial cell and in solution, contains two immunoglobulin-binding domains and two other domains that bind to the C3 complement component. The cellassociated Sbi has been demonstrated to be capable of preventing Fc-mediated protective functions such as opsonization and complement activation, by blocking complement, or Fc receptor binding, to the IgG constant regions. In addition, the portion of the Sbi protein that is released into the extracellular fluid has also been shown to be protective for the bacterium. Although the precise nature of the protective mechanism is still being investigated, it is thought to result from the depletion of complement components in the vicinity of the bacterium. S. aureus has also evolved mechanisms that take advantage of host complement control strategies. A bacterial cell-surface protein, Clumping factor A (ClfA) is able to bind factor I, thus localizing it to the bacterial surface, where it cleaves and inactivates any C3b that has bound there (5). Another complement inhibitory protein secreted by S. aureus acts in a slightly different way. The Chemotaxis Inhibitory Protein of S. aureus (CHIPs) binds two chemotactic receptors on neutrophils, the C5a receptor, and are potentially able to discriminate between host and pathogen. Thus, C1q may have expressed the capacity to recognize foreign molecules at a very early point in time, independent of antibody binding and in a manner similar to that of MBL. Mannose-binding lectins have been identified in lampreys, thus placing the origin of the MBLs at least as far back as the early vertebrates. Indeed, functional lectin pathways have been characterized in ascidians. However, although genes encoding MBL-like proteins have been characterized in ascidian genomes, the nature of their relationship to vertebrate MBL proteins is still unclear. The next three gene families to be considered all encode proteins that are implicated in the cleavage of the C3, C4, and C5 components of complement. The first of these fami- the formylated peptide receptor, thus effectively eliminating the capacity of neutrophils to respond to C5a-mediated chemotactic signals. Other Staphylococcal proteins are active in the inhibition of neutrophil binding to the membrane surfaces of the endothelial cells lining blood capillaries, and thus retard the extravasation of neutrophils in the vicinity of a local infection. The variety and effectiveness of these inhibitory mechanisms that have evolved in just one strain of bacteria emphasize for us the importance of the complement system in the normal control of bacterial infections, and the complexity that researchers and clinicians face when attempting to develop the next generation of antibacterial therapies. Foster, T .J. 2005. Immune evasion by staphylococci. Nature Reviews Microbiology 3:948– 958. Laarman, A., F. Milder, J. van Strijp, and S. Rooijakkers. 2010. Complement inhibition by gram-positive pathogens: Molecular mechanisms and therapeutic implications. Journal of Molecular Medicine (Berlin) 88:115–120. Serruto, D., R. Rappuoli, M. Scarselli, P. Gros, and J. A. van Strijp. 2010. Molecular mechanisms of complement evasion: Learning from staphylococci and meningococci. Nature Reviews Microbiology 8:393–399. lies is composed of factor B (Bf) and the serine protease C2; the second family comprises the serine proteases MASP-1, MASP-2, MASP-3, C1r, and C1s; and the third family contains the C3 family members C3, C4, and C5, which are not themselves proteases, but which share a common internal thioester and are subject to protease cleavage and activation. Studies of a number of invertebrate species have indicated that the genome of each species contains only a single copy representative of each of these three gene families. In contrast, one or more gene duplication events have occurred within each gene family in most vertebrates, suggesting that the gene duplications found in vertebrates probably occurred in the early stages of jawed vertebrate evolution. Genes for each of the Bf, C3, and MASP families have been identified from all 218 PA R T I I | Innate Immunity Alternative pathway Lectin pathway Classical pathway C3/C4/C5 Bf/C2 MASP/C1r,s TCC (MAC) Mammalia Reptilia/Aves Amphibia Teleostei Chondrichthyes Deuterostomes Agnatha Urochordata Cephalochordata Hemichordata Echinodermata Protstomia Cnidaria 1000 500 Million years ago 0 FIGURE 6-17 Evolution of complement components. The appearance of each of the three major complement pathways is illustrated by gray arrows, and the timing of the gene duplications that gave rise to the classical complement pathway is indicated by double-headed arrows. [Adapted from Nonaka, M., and A. Kimura, 2006, Genomic view of the evolution of the complement system, Immunogenetics 58:701.] TABLE 6-6 Complement system pathways in the major groups of deuterostome animals Animal group Alternate pathways Classical pathway Lectin pathway TCC complex Antibodies present? Mammals ⫹ ⫹ ⫹ ⫹ ⫹ Birds ⫹ ⫹ ⫹ ⫹ ⫹ Reptiles ⫹ ⫹ ⫹ ⫹ ⫹ Amphibians ⫹ ⫹ ⫹ ⫹ ⫹ Teleost fish ⫹ ⫹ ⫹ ⫹ ⫹ Cartilaginous fish ⫹ ⫹ ⫹ ⫹ ⫹ Agnathan fish ⫹  ⫹ ?  Tunicates ⫹  ⫹ ?  Echinoderms ⫹  ? ?  After J. O. Sunyer et al., 2003. Evolution of complement as an effector system in innate and adaptive immunity. Immunologic Research 27:549–564, Figure 2. The Complement System invertebrate deuterostomes that have so far been analyzed, with the single exception of the MASP family gene, which is missing from the sea urchin (Phylum Echinodermata) genome. The situation in protostomes is somewhat more complicated. Whereas members of the C3 and Bf families have been identified in some very early protostomes—for example, the horseshoe crab and the carpet-shell clam (Phyla: Arthropoda and Mollusca respectively)—analysis of whole genome sequences of insects such as Drosophila melanogaster (Arthropoda) or the nematode Caenorhabditis elegans (Nematoda) have failed to locate any proto-complement protein sequences. This contrasts starkly with the presence of all three family genes in sea anemones and corals (Phylum, Cnidaria), and suggests that the ability to encode proteins of the complement system has been lost from the genome of the common, model-system protostomes D. melanogaster and C. elegans. The absence of genes encoding any of the later, cytolytic components of the complement cascade in protostomes and cnidarians supports the hypothesis that the proto-complement systems of cnidarians and protostomes function by opsonization. The proteins encoded by the fifth gene family, C6, C7, C8, C8, and C9, together make up the terminal complement complex (TCC). They share a unique domain structure that enables molecular phylogenetic analysis. In particular, the mammalian TCCs, C6, C7, C8, C8, and C9 all share the MAC/perforin (MACPF) domain, in addition to other domains common among two or more members of this protein group. Mammals, birds, and amphibians share all TCC genes (with the exception of the bird C9 gene, which is miss- | CHAPTER 6 219 ing from the chicken genome sequence). Although the full complement of MAC proteins has not yet been documented in cartilaginous fishes such as the sharks (the earliest known organisms to possess an adaptive immune system), a gene encoding a C8 subunit orthologous to mammalian C8 has been cloned and characterized. Furthermore, the serum of sharks has been known for decades to express hemolytic activity, and microscopic analysis of the pores formed on target cell surfaces reveals a transmembrane pore structure indistinguishable from that induced by a MAC attack. It thus seems likely that the functionality of the MAC emerged not long after the adaptive immune system. Genomic analysis has traced the origin of the TCC genes back to before the divergence of the urochordates, cephalochordates, and vertebrates (see Figure 6-17) as ascidians (urochordates) and amphioxus (cephalochordates) possess primitive copies of TCC genes. However, the ascidian and amphioxus TCC proteins could not have functioned in a manner similar to that of later vertebrate species because they lack the domain responsible for interacting with the C5 protein. Intriguingly, toxins of the venomous sea anemone, which express a very high hemolytic potential, are closest in structure to the TCCs of the complement pathway. In summary, complement evolution is based on the diversification and successive duplications of members of five gene families that evolved first in response to the need for microbial recognition and opsonization and then responded to the appearance of the adaptive immune system by additional gene duplication and functional adaptation to form the complement system as we know it today. S U M M A R Y ■ ■ ■ ■ ■ ■ The complement system comprises a group of serum proteins, many of which circulate in inactive forms that must first be cleaved or undergo conformational changes prior to activation. Complement proteins include initiator molecules, enzymatic mediators, membrane-binding components or opsonins, inflammatory mediators, membrane attack proteins, complement receptor proteins, and regulatory components. Complement activation occurs by three pathways—classical, lectin, and alternative—which converge in a common sequence of events leading to membrane lysis. The classical pathway is initiated by antibodies of the IgM or IgG classes binding to a multivalent antigen. This allows binding of the first component of complement, C1q, and begins the process of complement deposition. The lectin pathway is initiated by binding of lectins such as mannose-binding lectin or members of the ficolin family to microbial surface carbohydrates. The alternative pathway is initiated when the third component of complement, C3, undergoes nonspecific break- down. C3b is formed and adheres to a cell surface. Inadvertent initiation is prevented by the presence of control proteins on host cell membranes. ■ ■ ■ ■ ■ The end result of the initiating sequence of all three pathways is the generation of enzymes that cleave C3 into C3a and C3b, and C5 into C5a and C5b. C3b opsonizes microbial cells and immune complexes, rendering them suitable for phagocytosis. Activation of the terminal components of the complement cascade C5b, C6, C7, C8 and C9 results in the deposition of a membrane attack complex (MAC) onto the microbial cell membrane. This complex introduces large pores in the membrane, preventing it from maintaining osmotic integrity and resulting in the death of the cell. Binding of the complement component C1q to apoptotic cells, apoptotic bodies, and immune complexes results in their opsonization with C3b and subsequent phagocytosis. Patients deficient in the early complement components do not clear apoptotic cells efficiently and suffer disproportionally from autoimmune disease such as SLE. 220 ■ ■ ■ PA R T I I | Innate Immunity In addition to acting during host defense against infection, complement proteins also bind to receptors on the surfaces of antigen-presenting cells, T cells, and B cells, inducing interleukin production and aiding in their activation. A system of regulatory complement proteins ensures that inadvertent complement activation on the surface of host cells is prevented and controlled, by deactivating complement components on the surface of host cells, and ensuring that regulatory proteins are bound specifically to host, but not to microbial cell membranes. Patients suffering from complement deficiencies often present with immune complex disorders and suffer disproportionally from infections by encapsulated bacteria such as Neisseria meningitidis. Animal models exist for most complement deficiencies. ■ ■ Underscoring the importance of complement in the immune system, a broad variety of complement evasion strategies has evolved in viruses, bacteria, fungi, and parasites, including mimicking regulatory proteins, interfering with the interactions between antibodies and complement components, or between proteins of the complement pathways, or by destruction of the complement components. The genes encoding complement components belong to five families. Genes of the alternative pathway components appear first in evolution, and those encoding the terminal complement components appear last. Thus, prior to the emergence of adaptive immunity, complement served its protective functions by mediating phagocytosis. R E F E R E N C E S Alexander, J. J., A. J. Anderson, S. R. Barnum, B. Stevens, B., and A. J. Tenner. 2008. The complement cascade: Yin-Yang in neuroinflammation—neuro-protection and -degeneration. Journal of Neurochemistry 107:1169–1187. Al-Sharif, W. Z., J. O. Sunyer, J. D. Lambris, and L. C. Smith. 1998. Sea urchin coelomocytes specifically express a homologue of the complement component C3. Journal of Immunology 160:2983–2997. Azumi, K., et al. 2003. Genomic analysis of immunity in a Urochordate and the emergence of the vertebrate immune system: “Waiting for Godot.” Immunogenetics 55:570–581. Behnsen, J., et al. 2010. Secreted Aspergillus fumigatus protease Alp1 degrades human complement proteins C3, C4, and C5. Infection and Immunity 78:3585–3594. Blom, A. M., T. Hallstrom, and K. Riesbeck, K. 2009. Complement evasion strategies of pathogens—acquisition of inhibitors and beyond. Molecular Immunology 46:2808–2817. Botto, M., et al. 1998. Homozygous C1q deficiency causes glomerulonephritis associated with multiple apoptotic bodies. Nature Genetics 19:56–59. Kang, Y. S., et al. 2006. A dominant complement fixation pathway for pneumococcal polysaccharides initiated by SIGN-R1 interacting with C1q. Cell 125:47–58. Kemper, C., and J. P. Atkinson. 2007. T-cell regulation: With complements from innate immunity. Nature Reviews Immunology 7:9–18. Kemper, C., and D. E. Hourcade. 2008. Properdin: New roles in pattern recognition and target clearance. Molecular Immunology 45:4048–4056. Lambris, J. D., D. Ricklin, and B. V. Geisbrecht. 2008. Complement evasion by human pathogens. Nature Reviews Microbiology 6:132–142. Litvack, M. L. and N. Palaniyar. 2010. Soluble innate immune pattern-recognition proteins for clearing dying cells and cellular components: Implications on exacerbating or resolving inflammation. Innate Immunity 16:191–200. Longhi, M. P., C. L. Harris, B. P. Morgan, and A. Gallimore, A. 2006. Holding T cells in check: A new role for complement regulators? Trends in Immunology 27:102–108. Dunkelberger, J. R., and W. C. Song. 2010. Complement and its role in innate and adaptive immune responses. Cell Research 20:34–50. Markiewski, M. M., B. Nilsson, K. N. Ekdahl, T. E. Mollnes, and J. D. Lambris. 2007. Complement and coagulation: Strangers or partners in crime? Trends in Immunology 28:184–192. Elward, K., et al. 2005. CD46 plays a key role in tailoring innate immune recognition of apoptotic and necrotic cells. Journal of Biological Chemistry 280:36342–36354. Miller, D. J., et al. 2007. The innate immune repertoire in Cnidaria: Ancestral complexity and stochastic gene loss. Genome Biology 8:R59. Fearon, D. T., and M. C. Carroll. 2000. Regulation of B lymphocyte responses to foreign and self-antigens by the CD19/CD21 complex. Annual Review of Immunology 18:393–422. Nonaka, M., and A. Kimura. 2006. Genomic view of the evolution of the complement system. Immunogenetics 58:701–713. Gros, P., F. J. Milder, and B. J. Janssen. 2008. Complement driven by conformational changes. Nature Reviews Immunology 8:48–58. Price, J. D., et al. 2005. Induction of a regulatory phenotype in human CD4⫹ T cells by streptococcal M protein. Journal of Immunology 175:677–684. Jensen, J. A., E. Festa, D. S. Smith, and M. Cayer. 1981. The complement system of the nurse shark: hemolytic and comparative characteristics. Science 214:566–569. Ricklin, D., and J. D. Lambris. 2008. Compstatin: A complement inhibitor on its way to clinical application. Advances in Experimental and Medical Biology 632:273–292. The Complement System | CHAPTER 6 221 Roozendaal, R., and M. C. Carroll. 2007. Complement receptors CD21 and CD35 in humoral immunity. Immunological Reviews 219:157–166. Zipfel, P. F., and C. Skerka. 2009. Complement regulators and inhibitory proteins. Nature Reviews Immunology 9:729–740. Rosen, A. M., and B. Stevens. 2010. The role of the classical complement cascade in synapse loss during development and glaucoma. Advances in Experimental and Medical Biology 703:75–93. Useful Web Sites Stevens, B., et al. 2007. The classical complement cascade mediates CNS synapse elimination. Cell 131:1164–1178. Sunyer, J. O., et al. 2003. Evolution of complement as an effector system in innate and adaptive immunity. Immunologic Research 27:549–564. Suzuki, M. M., N. Satoh, and M. Nonaka. 2002. C6-like and C3-like molecules from the cephalochordate, amphioxus, suggest a cytolytic complement system in invertebrates. Journal of Molecular Evolution 54:671–679. Ward, P. A. 2009. Functions of C5a receptors. Journal of Molecular Medicine 87:375–378. Zhu, Y., S. Thangamani, B. Ho, and J. L. Ding. 2005. The ancient origin of the complement system. European Molecular Biology Organization Journal 24:382–394. S T U D Y www.complement-genetics.uni-mainz.de The Complement Genetics Homepage from the University of Mainz gives chromosomal locations and information on genetic deficiencies of complement proteins. www.cehs.siu.edu/fix/medmicro/cfix.htm A clever graphic representation of the basic assay for complement activity using red blood cell lysis, from D. Fix at the University of Southern Illinois, Carbondale. www.youtube.com/watch?vⴝy2ep6j5kHUc www.youtube.com/watch?vⴝAIjaiJV4m2g There are a number of animations of the complement cascade available on the Internet. These links are to clear animations. Q U E S T I O N S 1. Indicate whether each of the following statements is true or 5. Complement activation can occur via the classical, alterna- false. If you think a statement is false, explain why. tive, or lectin pathway. a. A single molecule of bound IgM can activate the C1q a. How do the three pathways differ in the substances that b. c. d. e. f. component of the classical complement pathway. The enzymes that cleave C3 and C4 are referred to as convertases. C3a and C3b are fragments of C3 that are generated by proteolytic cleavage mediated by two different enzyme complexes. Nucleated cells tend to be more resistant to complement-mediated lysis than red blood cells. Enveloped viruses cannot be lysed by complement because their outer envelopes are resistant to pore formation by the membrane attack complex (MAC). MBL has a function in the lectin pathway analogous to that of IgM in the classical pathway, and MASP-1 and MASP-2 take on functions analogous to C1 components. 2. Explain why serum IgM cannot activate complement prior to antigen binding. 3. Genetic deficiencies have been described in patients for all of the complement components except factor B. Particularly severe consequences result from a deficiency in C3. Describe the consequences of an absence of C3 for each of the following: a. Initiation of the classical and alternate pathways b. Clearance of immune complexes c. Phagocytosis of infectious bacteria 4. Describe three ways in which complement acts to protect the host during an infection. can initiate activation? b. Which parts of the overall activation sequence differ among the three pathways, and which parts are similar? c. How does the host ensure that inadvertent activation of the alternative pathway on its own healthy cells does not lead to autoimmune destruction? 6. Briefly explain the mechanism of action of the following complement regulatory proteins. Indicate which pathway(s) each protein regulates. a. b. c. d. e. f. C1 inhibitor (C1INH) C4b-binding protein (C4bBP) Decay-accelerating factor (DAF) Membrane cofactor of proteolysis protein, MCP (CD46) Protectin (CD59) Carboxypeptidase N 7. Explain why complement deficiencies in the early compo- nents of complement give rise to immune-complexmediated disorders such as systemic lupus erythematosus. 8. You have prepared knockout mice with mutations in the genes that encode various complement components. Each knockout strain cannot express one of the complement components listed across the top of the table below. Predict the effect of each mutation on the steps in complement activation and on the complement effector functions indicated in the table using the following symbols: NE no effect; D process/function decreased but not abolished; A process/function abolished. PA R T I I 222 | Innate Immunity Complement component knocked out C1q C4 C3 C5 C9 Factor B MASP-2 Formation of classical pathway C3 convertase Formation of alternative pathway C3 convertase Formation of classical pathway C5 convertase Formation of lectin pathway C3 convertase C3b-mediated opsonization Neutrophil chemotaxis and inflammation Cell lysis As shown in the figure below, two flow cytometric histograms of red blood cells were obtained from a patient, stained with antibodies toward CD59 (Protectin) (part A) and CD55 (Decay Accelerating Factor or DAF) (part B). On the right of each histogram is a large population of cells expressing relatively high levels of each antigen (population I). Populations II express midrange levels of the two antigens, and populations III express levels of antigen below statistical detectability. (Detecting laser voltages on flow cytometers are normally set such that negative staining populations show levels of fluorescence below 101 on the lower logarithmic scale. Note also the creep of cells up the 100 axis on each plot.) a. Can you offer an explanation as to why this patient may b. Why do you think a single patient can generate red blood have red blood cells expressing low levels of these two particular antigens? From what disease do you think this patient is suffering? c. Now look at parts C and D of the same figure, which CLINICAL FOCUS QUESTION cells expressing three different levels of CD59 and CD55? display the expression of two other markers on red The Complement System blood cells from the same patient. In these flow cytometric profiles, the investigators have shown you quadrant markers to indicate whether the cells they are investigating are considered to be positive or negative for the proteins labeled on the axes. For example, the cell population high and right in part C is considered to express both CD66abce and CD16, whereas the cell population low and left is negative for both of these markers. Using your answer for part (a) as a starting point, what can you deduce about the biochemistry of the membrane proteins CD16, CD66abce, CD46, and CD14? CHAPTER 6 a. In the figure below, a flow cytometric histogram describes the numbers of human Jurkat T cells expressing low and high levels of CD46 after treatment with an apoptosis-inducing reagent. On the figure, label the cell population that you think is undergoing apoptosis and explain your reasoning. 200 CD 46 Counts 160 120 80 40 0 100 101 102 FL2-H 103 104 223 On the flow cytometric dot plot shown below, indicate the following and explain your labeling: a. Where you would expect to find a healthy T-cell population? b. Where you would expect to find a T-cell population undergoing apoptosis? C1q ANALYZE THE DATA | CD46 This page lelt intentionally blank. Site of DNA cleavage The Organization and Expression of Lymphocyte Receptor Genes 3' T o protect its host, the immune system must recognize a vast array of rapidly evolving microorganisms. To accomplish this, it must generate a diverse and flexible repertoire of receptor molecules, while minimizing the expression of receptors that recognize self antigens. As described in Chapter 3, each B or T lymphocyte expresses a unique antigen-specific receptor. When these receptors bind to their corresponding antigens under the appropriate conditions (described in Chapters 11 and 12), T and B lymphocytes proliferate and differentiate into effector cells that eliminate the microbial threat (Chapter 13). In Chapter 3, we described the biochemistry of the T and B lymphocyte receptors and the secreted antibodies formed by B lymphocytes following antigen stimulation. We also outlined the experiments which demonstrated that secreted antibodies are identical in antigen-binding specificity to the B-cell receptors of the secreting cell. In this chapter, we address the question of how an organism can encode and express receptors capable of recognizing a constantly evolving universe of microbial threats using a finite amount of genetic information. The production of specific lymphocyte receptors employs a number of genetic mechanisms that are unique to the immune system. In 1987, Susumu Tonegawa won the Nobel Prize for Physiology or Medicine “for his discovery of the genetic principle for generation of antibody diversity,” a discovery that challenged the fundamental concept that one gene encoded one polypeptide chain. Tonegawa and his colleagues showed that the antibody light chain was encoded in the germ line by not one but three families of gene segments separated by kilobases of DNA (Figure 7-1). (The germ-line DNA is the genetic information encoded in the sperm and egg, which can be passed on to future generations.) Their work demonstrated that two DNA segments, one from each family, are conjoined, only in B lymphocytes, to create the mature form of the light-chain variable region of the immunoglobulin (Ig) gene. A third segment encodes the constant region of the gene. Different combinations of RSS Antigen receptor gene segment 5' 7 12 bp ACAAAAACC–3' 5 '– C A C A G T G 23 bp Heptamer Spacer Nonamer The Recombination Signal Sequence (RSS) serves as the site of DNA cleavage. The RSS is composed of conserved heptamer and nonamer sequences, separated by a spacer region of 12 or 23 bp, which is conserved in length, but not in sequence. Cleavage occurs at the junction of the heptamer and the variable region coding segment. [Adapted from D. G. Schatz and Y. Ji, 2011, Recombination centres and the orchestration of V(D)J recombination, Nature Reviews. Immunology 11:251–263.] ■ The Puzzle of Immunoglobulin Gene Structure ■ Multigene Organization of Ig Genes ■ The Mechanism of V(D)J Recombination ■ B-Cell Receptor Expression ■ T-Cell Receptor Genes and Expression segments are used in each B cell, to create the diverse repertoire of light-chain receptor genes. Subsequent experiments have shown that all of the B- and T-cell receptor genes are assembled from multiple gene segments by similar rearrangements. We describe below the unique genetic arrangements of T- and B-cell receptor gene segments, and the mechanisms by which they are rearranged and expressed. We will address here only those mechanisms that shape the receptor repertoires of mouse and human naïve B and T cells, which have not yet been exposed to antigen. 225 PA R T I I I 226 5' V V | Adaptive Immunity: Antigen Receptors and MHC V 3' J 23 kb 5' 3' V J C 2.5 kb C (a) There are many variable regions, but just a few constant regions. V region (thousands, heavy and light chain) C region (four, light chain; or eight, heavy chain) (b) The same V region can be found connected to different C regions. V region C region FIGURE 7-1 The antibody light-chain gene encodes three families of DNA segments. During B-cell development, one V segment and one J segment (which encode contiguous parts of the light-chain variable region) join together with the C (constant) region to form the gene for the antibody light chain. This gene rearrangement occurs in the DNA, prior to gene transcription into mRNA. Additional layers of diversity are generated in B cells following antigen exposure and those will be addressed in Chapter 12. These powerful mechanisms include the generation of antibodies of different heavy-chain classes, each capable of mediating a discrete set of effector functions, as well as the creation of modified antigenbinding regions by somatic hypermutation, followed by antigen-driven selection. Both of these processes are triggered in human and mouse B cells only after antigen contact. The Puzzle of Immunoglobulin Gene Structure The immune system relies on a vast array of B-cell receptors (BCRs) that possess the ability to bind specifically to a correspondingly large number of potential pathogens. The first indication of the immense size of the antibody repertoire was provided by immunologists using synthetic molecules to stimulate antibody production. They discovered that antibodies can discriminate between small synthetic molecules differing in as little as the position of an amino or hydroxyl group on a phenyl ring. If the immune system can discriminate between small molecules that it had presumably never encountered during evolutionary selection and that differ in such subtle ways, then, it was reasoned, the number of potential antibodies must be very large indeed. A series of experiments conducted in the late 1970s and early 1980s estimated the number of different BCRs generated in a normal mouse immune system to be at least 107, but we now know that estimate was many orders of magnitude too small. Investigators trying to make sense of Ig genetics were also faced with an additional puzzle: protein sequencing of mouse and human antibody heavy and light chains revealed that the first (amino terminal) 110 amino acids of antibody heavy and light chains are extremely variable among different antibody molecules. This region was therefore defined as the variable (V) region of the antibody molecule. In contrast, the remainder of the light and heavy chains could be classified into one FIGURE 7-2 Early sequencing studies indicated that the light chain may be encoded in more than one segment. (a) Many Ig variable regions in both heavy and light chains could be found in association with few constant regions. (b) The same variable region can be found in contiguity with several different heavy-chain constant regions. of only four sequences (light chain) or eight sequences (heavy chain) (Figure 7-2a) and was, therefore, named the constant (C) region. This raised an intriguing question: If each of 105 to 107 antibodies is encoded by a separate gene, how could the constant region part of the gene remain constant in sequence in the face of evolutionary drift? Furthermore, antibodies could be found in which the same antibody variable (V) region was associated with more than one heavy-chain constant (C) region (Figure 7-2b), lending further support to the possibility that the expression of the variable and constant regions of each antibody chain were independently controlled. (These additional constant regions are generated by the antigen-induced process of class switch recombination, discussed in Chapter 12. In the current context, the only point of relevance is the independence with which the variable and constant regions appear to be expressed.) It was rapidly becoming clear that the solution to the antibody gene puzzle was more complex than had previously been imagined. Investigators Proposed Two Early Theoretical Models of Antibody Genetics Classical germ-line theories of antibody diversity suggested that the genetic information for each antibody is encoded, in its entirety, within the germ-line genome. This means that the genes encoding the entire sequence of every antibody heavy or light chain that the animal could ever make must be present in every cell. However, a quick calculation is sufficient to demonstrate that if there are 107 or more antibodies, each of which requires approximately 2,000 nucleotides, this would require massive expenditure of genetic information— indeed, more DNA would be required to encode the receptors of the immune system than is available to the organism. Although arguments were initially made that the dedication of a considerable fraction of the genome to the immune system may represent a reasonable evolutionary strategy, it The Organization and Expression of Lymphocyte Receptor Genes became clear, as estimates of the size of the antibody repertoire were revised upward, that there simply was not enough DNA to go around and new ideas must be found. In 1965, William Dreyer and J. Claude Bennett proposed that antibody heavy and light chains are each encoded in two separate segments in the germ-line genome and that one of each of the V region- and C region-encoding segments are brought together in B-cell DNA to form complete antibody heavy and light-chain genes. The idea that DNA in somatic cells might engage in recombinatorial activity was revolutionary. However, germ-line theorists began to modify their ideas to embrace this possibility. Others suggested the equally innovative idea that the number of variable region genes in the genome might be extremely limited, and proposed the somatic hypermutation theory. According to this hypothesis, a limited number of antibody genes is acted upon by unknown mutational mechanisms in somatic cells to generate a diverse receptor repertoire in mature B lymphocytes. This latter idea had the advantage of explaining how a large repertoire of antibodies could be generated from a relatively small number of genes, but the disadvantage that such a process, like the somatic cell gene recombination suggested by others, had never been observed. There was additional argument over whether mutation would occur prior to, or only after, antigen contact and, therefore, whether mutation was responsible for generation of the so-called “primary repertoire” that exists prior to antigen binding by the B cell. | CHAPTER 7 227 Heated debate continued between the proponents of modified germ-line versus somatic mutation theories throughout the early 1970s, until a seminal set of experiments revealed that both sides were correct. We now know that each antibody molecule is encoded by multiple, germline, variable-region gene segments, which are rearranged differently in each naïve immune cell to produce a diverse primary receptor repertoire. These rearranged genes are then further acted upon after antigen encounter by somatic hypermutation and antigenic selection, resulting in an expanded and exquisitely honed repertoire of antigenspecific B cells (see Chapter 12). Breakthrough Experiments Revealed That Multiple Gene Segments Encode the Light Chain From the mid 1970s until the mid 1980s, a small group of brilliant immunologists completed a series of experiments that fundamentally altered the way in which scientists think about the genetics of immune receptor molecules. The first breakthrough occurred when Susumu Tonegawa showed, as Dreyer and Bennett had predicted, that multiple gene segments encode the antibody light chain. His achievement is all the more impressive because the modern tools of molecular biology were not yet available. The experiment he performed with Nobumichi Hozumi is described in Classic Experiment Box 7-1. BOX 7-1 CLASSIC EXPERIMENT Hozumi and Tonegawa’s Experiment: DNA Recombination Occurs in immunoglobulin Genes in Somatic Cells The paradigm-shifting experiment of Hozumi and Tonegawa was designed to determine whether the DNA encoding Ig light-chain constant and variable regions existed in separate segments in non antibody-producing cells, such that a single constant region gene could associate with different variable region genes in different B cells. They hypothesized that this might be the case, because of Ig amino acid sequencing data showing that the sequences of the constant regions of many Ig light chains were identical. They reasoned that if multiple copies of a constant region gene existed, then each copy would be expected to accumulate silent or neutral mutations over time. The most likely explanation for the finding of a single con- stant region amino acid sequence arranged in tandem with many different variable regions was, therefore, that a single constant region gene segment was cooperating with multiple variable region gene segments in different cells or situations to generate the light-chain gene. However, this notion was heretical to a generation of scientists brought up on the concept that one gene encodes one polypeptide chain. Scientists are accustomed to thinking about the concept of alternative RNA splicing, wherein different proteins may be encoded by the same piece of DNA by differentially using particular RNA segments, cut and spliced following transcription of a long precursor RNA transcript. However, this experiment asked an entirely new question: Can a piece of DNA change its place on a chromosome in a somatic cell? To serve as their source of germ-line DNA, Hozumi and Tonegawa used DNA from an organ in which Ig genes would not be expressed, embryonic liver in this case. (Sperm and egg DNA would have been much more difficult to obtain.) For B-cell DNA, they used DNA from an antibodyproducing plasma cell tumor line, MOPC 321, which secretes fully functional  light chains. They separately cut both sets of DNA with the same restriction enzyme and used radioactive probes to determine the sizes of the DNA fragments on which the variable and constant regions of the light-chain gene were found in the two sets of cells. (continued) 228 PA R T I I I CLASSIC EXPERIMENT | Adaptive Immunity: Antigen Receptors and MHC (continued) Below, we describe how Hozumi and Tonegawa conducted their experiment, and indicate how the experimental protocol would be modified using the reagents and techniques available to a modern molecular biologist. We then display their data as it would appear both in the modern gel format as well as in its original form. Hozumi and Tonegawa: a. Purified genomic DNA from embryonic liver cells and from the MOPC 321 tumor cell line. b. Cut the two sets of genomic DNA with a restriction endonuclease (BamHI) and separated the DNA fragments using electrophoresis. (Today we would use a polyacrylamide gel and electrophorese submicrogram quantities of samples over the course of a few hours; Hozumi and Tonegawa used a footlong agarose gel that needed 2 liters of agarose. They loaded 5 mg of DNA and ran the gel for 3 days.) c. Made 125I-labeled nucleic acid probes specific for the two regions of mouse  light chain. One of Tonegawa’s probes was a full-length radiolabeled piece of mRNA encoding the entire -chain sequence. The other probe was the 3 half of the sequence, which would hybridize to the constant, but not to the variable region of the -chain gene. Today, making enzymatically labeled, stable DNA probes for any particular sequence is a safe and relatively straightforward task. In Hozumi and Tonegawa’s day, this task was significantly more challenging. d. Probed the nuclease-generated DNA fragments to determine the size of the fragments carrying the variable and constant region sequences. We would normally use a Southern blot procedure, electrophoresing the fragments, blotting the gel with nitrocellulose paper, and then probing the paper with enzyme-labeled fragments complementary to the sequences of interest, developing the blots with Antibody-producing tumor cells from same strain of mouse Embryonic liver cells (Germline DNA) Extract DNA from cells Cut DNA with Restriction Endonuclease Gel electrophoresis 1 2 3 4 Blot onto nitrocellulose paper and probe with radioactive sequences complementary to antibody variable and constant regions Germline DNA C-region probe 1 DNA from antibody producing cells Probe complementary to whole light chain 2 C-region probe 3 Probe complementary to whole light chain 4 FIGURE 1 The  light-chain gene is formed by DNA recombination between variable and constant region gene segments. DNA from embryonic liver cells was used as a source of germ-line DNA, and DNA from a B-cell tumor cell line was used as an example of DNA from antibody-producing cells. In embryonic liver, the DNA sequences encoding the variable and constant regions, respectively, were located on different restriction endonuclease fragments. However, these two sequences were colocated on a single restriction fragment in the myeloma DNA. (See text for details of the experiment.) luminescent or fluorescent substrates. Hozumi and Tonegawa’s approach was much more timeconsuming. They cut the gel into about 30 slices, melted the agarose, and separately eluted the DNA from each slice. To each DNA sample, they added radiolabeled RNA, allowed it to anneal, and removed the un-annealed RNA using RNase. The radioactivity remaining in each fraction was then plotted against the size of the DNA in the slice. Figure 1 shows the results that would be obtained from a modern-day Southern The Organization and Expression of Lymphocyte Receptor Genes | CHAPTER 7 229 BOX 7-1 blot of Tonegawa’s fragments. The germline DNA blot probed with the constant region mRNA sequence probe (lane 1) shows only a single band. This indicates that the restriction endonuclease is not cutting in the middle of the light-chain constant region, but rather that the whole constant region sequence is encoded within a single fragment. In contrast, probing the blot with the whole light-chain sequence (V and C, lane 2) yields two bands. Since the difference between lanes 1 and 2 is the presence of the variable region sequence in the probe used for lane 2, the presence of two bands in lane 2 indicates that the restriction endonuclease used to generate the DNA fragments is cutting somewhere in between the variable and constant regions in the germ-line DNA, such that the variable region lies on a DNA fragment distinct from that bearing the constant region. Analyzing the blots from the plasma cell tumor, we first note that the large fragment containing the constant region DNA and the midsized fragment bearing the variable region DNA have disappeared. Both the variable and constant region gene segments are now located on smaller fragments. This implies that a new pair of restriction endonuclease sites now forms the boundaries of each of the constant and variable region fragments. Right away, we can tell that the DNA environment around the light-chain genes changes as the B cell differentiates. Next, we note that the sizes of the DNA fragments on which the constant and variable region gene segments are located are apparently the same. This implies, although it does not yet prove, that the movement of the constant and/ or variable region gene segments has brought them into close proximity with one another, such that the constant and variable region gene segments co-locate on the same fragment. An alternative explanation is that they have both altered their locations and the similarity in the size of the fragments is coincidental. DNA sequencing supported the former FIGURE 2 The original data from Hozumi and Tonegawa’s classic experiment proving immunoglobulin gene recombination occurs in B cells. DNA from the sources shown was run out on an agarose gel for 3 days at 4C. The gel was sliced, the DNA eluted, and then each sample was hybridized with radiolabeled probes for either the constant region of the  light chain, or the whole chain. The plot shows the amount of radioactivity in each fraction as a function of the migration distance, which reflects the molecular weight of the DNA fragment. (See text for details.) [ From N. Hozumi and S. Tonegawa, 1976, Evidence for somatic rearrangement of immunoglobulin genes coding for variable and constant regions, Proceedings of the National Academy of Sciences USA 73:3628–3632.] interpretation: as the B cell differentiates, the variable and constant region gene segments are moved from distant regions of the chromosome into close apposition with one another. It is all too easy, with modern molecular biology technologies, to forget what a tour de force this experiment actually represented. Few of the reagents or pieces of apparatus we commonly encounter in the molecular biology laboratory today were available to Hozumi and Tonegawa— they had to make their own. The original paper speaks of purifying their own restriction endonuclease (BamHI) from a bacterial sample obtained from a colleague, and of performing the electrophoresis at 4oC for 3 days. Furthermore, technical difficulties precluded their being able to make a good 5 (V-region) probe. They therefore probed each sample with a 3 (C-region) probe, as well as one that bound to the whole light chain, and then inferred 5 (V-region) binding by subtraction. It was an extraordinary piece of work. Their original data are shown in Figure 2. What happened to the DNA on the alternative allele that did not encode the tumor cell secreted light chain? Subsequent analysis of the DNA from this tumor showed that the DNA from both chromosomes had undergone rearrangement. Hozumi and Tonegawa were fortunate that the variable regions used by both rearrangements were close to one another and so the fragment patterns overlapped. Hozumi, N., and S. Tonegawa. 1976. Evidence for somatic rearrangement of Ig genes coding for variable and constant regions. Proceedings of the National Academy of Sciences USA 73:3628–3632. 230 PA R T I I I | Adaptive Immunity: Antigen Receptors and MHC Tonegawa’s experiment showed that, in the DNA from non-antibody-producing, embryonic liver cells, there is a BamHI endonuclease site between the variable and constant regions. We know this because probes for the variable and constant regions each recognized a different DNA fragment in BamHI-digested germ-line DNA. However, in the antibodyproducing tumor cell, the DNA encoding the variable and constant regions appeared to be combined in just one fragment, thus demonstrating that DNA rearrangement must have occurred during the formation of an antibody lightchain gene (see Classic Experiment Box 7-1 for further details). Although this experiment demonstrated that the V and C regions of antibody genes were located in different contexts in the DNA of non-antibody-producing cells, it did not speak to their relative locations; indeed, the initial experiment did not rule out the possibility that the V and C fragments could be encoded on different chromosomes in the embryonic cells. However, sequencing experiments subsequently showed that the segments encoding the V and C genes of the  light chains are on the same chromosome and that, in non B cells, the two segments are separated by a long non-coding DNA sequence. The impact of this result on the biological community was profound. For the first time, DNA was shown to be cut and rejoined during the process of cell differentiation. This experiment not only provided the experimental proof of Dreyer and Bennett’s prediction; it also paved the way for the next surprising finding. Tonegawa’s experiment had identified the V and C segments encoding the kappa light chain. However, when scientists in his group sequenced the antibody light-chain DNA, they encountered another unexpected result. As expected, the embryonic (unrearranged) V region segment had, at its 5 terminus, a short hydrophobic leader sequence, a common feature of membrane proteins necessary to guide the nascent protein chain into the membrane (Figure 7-3a). A 93 bp sequence of non-coding DNA separated the leader sequence from a long stretch of DNA that encoded the first 97 amino acids of the V region. But the light-chain V region domain is approximately 110 amino acids long. Where was the coding information for the remaining 13 amino acids? Sequencing of the light-chain constant region fragment from embryonic DNA provided the answer. Upstream from the constant region coding sequence, and separated from it by a non-coding DNA segment of 1250 bp, were the 39 bp encoding the remaining 13 amino acids of the V region. This additional light-chain coding segment was named the joining (J) gene segment (shown in red in Figure 7-3a). Further sequencing of mouse and human light-chain variable and constant region genes confirmed Tonegawa’s second, astonishing finding. Not only are the variable and constant regions encoded in two separate segments, but the lightchain variable regions themselves are encoded in two separate gene segments, the V and J segments, that are made contiguous only in antibody-producing cells. The heavy-chain variable region gene was then shown to require yet a third genetic segment. Adopting a similar strategy of cloning and sequencing Ig heavy-chain genes, Lee Hood’s group identified a germ-line Heavy-chain Variable Region (VH) gene fragment that encoded amino acid residues 1–101 of the antibody heavy chain and a second fragment that included a Heavy-chain Joining gene segment (JH) that determined the sequence of amino acid residues 107–123. Neither of these segments contained the DNA sequence necessary to encode residues 102–106 of the heavy chain. Significantly, these missing residues were included in the third complementarity-determining region of the antibody heavy chain, CDR3, which provides contact residues in the binding of most antigens (see Chapter 3). Gene segments encoding this part of the antibody heavy chain were eventually located 5 of the J region in mouse embryonic DNA by Hood and colleagues (Figure 7-3b). The importance of the contribution made by this gene segment to the diversity of antibody specificities is denoted by its name; the Diversity (D) region. (Because there is no D region in the light chain, immunologists usually drop the subscript denoting the heavy chain.) Thus, the variable region of the heavy chain of the antibody molecule is encoded by three discrete gene segments, (a) Light chain V region gene segments in embryo (germline DNA) 5' Leader V J Downstream to C region gene segment 3' V segment (amino acids 1-97) J segment (amino acids 98-110) (b) Heavy chain V region gene segments in embryo (germline DNA) 5' Leader V D V segment (amino acids 1-101) D segment (amino acids 102-106 approx.) J 3' J segment (amino acids 107-123 approx.) Downstream to C region gene segments FIGURE 7-3 Antibody light chains are encoded in two segments—V and J, whereas antibody heavy chains are encoded in three segments—V, D, and J. | The Organization and Expression of Lymphocyte Receptor Genes TABLE 7-1 231 Kappa Light-Chain Genes Include V, J, and C Segments Chromosomal locations of immunoglobulin genes in humans and mice Human chromosome Mouse chromosome  light chain 22 16  light chain 2 6 Heavy chain 14 12 Gene CHAPTER 7 The mouse Ig locus spans 3.5 Mb and includes 91 potentially functional V genes, which have been grouped into 18 V gene families based on sequence homology (Figure 7-4a). (If two sequences are greater than 80% identical, they are classified as belonging to the same V gene family.) These V gene families can be further grouped into V gene clans, based again on sequence homology. Each V segment includes the leader exon encoding the signal peptide. Individual V segments are separated by non-coding gaps of 5 to 100 kb. The transcriptional orientation of particular V segments may be in the same or in the opposite direction as the constant region segment. The relative orientations of the variable and constant region segments do not affect the use of the segments, but do alter some details of the recombinational mechanism that creates the complete light-chain gene, as will be discussed later. Downstream of the V region cluster are four functional J segments and one pseudogene, or other nonfunctional open reading frame (Table 7-2). A similar arrangement is found in the human V locus, although the numbers of V and J gene segments vary. A single C segment is found downstream of the J region, and all kappa light-chain constant regions are encoded by this segment. and the variable region of the light chain by two segments, in the germ-line genome. These segments are brought together by a process of DNA recombination that occurs only in the B lymphocyte lineage to create the complete variable region gene. Furthermore, the DNA at the junction between the V and J segments for light chains, and at the VD and DJ junctions in heavy chains, accounts for the extraordinary diversity that was first observed by Kabat and Wu in the CDR3 regions of both heavy and light chains (see Chapter 3). Below, we will describe how further genetic diversity is generated at these junctions by additional processes unique to immune system genetics. Multigene Organization of Ig Genes Lambda Light-Chain Genes Pair Each J Segment with a Particular C Segment Recall that Ig proteins consist of two identical heavy chains and two identical light chains (see Chapter 3). The light chains can be either kappa () light chains or lambda () light chains. The heavy-chain, kappa, and lambda gene families are each encoded on separate chromosomes (Table 7-1). The mouse Ig locus spans a region of approximately 190 kb. Lambda light chains are only found in 5% of mouse antibodies because of a deletional event in the mouse genome that has eliminated most of the lambda light-chain variable region segments. It was therefore not surprising to discover that (a) κ-chain DNA 5' 3' L Vκ1 L Vκ(n) L Vκ2 Cκ Jκ ψ 23 kb 2.5 kb (b) λ-chain DNA 5' 3' L Vλ2 L Vλ3 Jλ2 Cλ2 Jλ4 Jλ3 L Vλ1 Cλ4 Cλ3 Jλ1 Cλ1 ψ 19 kb 55 kb 1.2 kb 2.0 1.3 kb kb 19 kb 1.4 kb 1.7 1.3 kb kb Cγ3 Cγ 1 (c) Heavy-chain DNA 5' 3' L VH1 L VH(n) DH1 DH14 JH1 JH4 6.5 kb Cμ Cδ 4.5 kb 55 kb 34 kb Cγ2b 21 kb Cγ2a 15 kb Cε 14 kb Cα 12 kb FIGURE 7-4 Organization of immunoglobulin germ-line gene segments in the mouse. The (a)  light chain and (b)  light chains are encoded by V, J, and C gene segments. The (c) heavy chain is encoded by V, D, J, and C gene segments. The distances in kilobases (kb) separating the various gene segments in mouse germ-line DNA are shown below each diagram. 232 TABLE 7-2 PA R T I I I | Adaptive Immunity: Antigen Receptors and MHC Immunoglobulin variable region gene numbers in humans and mice Human Mouse VK 34–48 functional; 8 ORFs; 30 pseudogenes. 91 functional; 9 ORFs; 60 pseudogenes JK 5 functional; multiple alleles 4 functional; 1 ORF V 33 functional; 6 ORFs; 36 pseudogenes 3–8 functional J 5 functional; 2 ORFs 4 functional; 1 pseudogene VH 38–44 functional; 4 ORFs; 79 pseudogenes 101 functional; 69 pseudogenes* D 23 functional; 4 ORFs 19 functional; 7 ORFs; 6 pseudogenes JH 6 functional; 3 pseudogenes; several different alleles found 4 functional; again, multiple haplotypes * Gene numbers of the mouse VH locus refer to the sequenced chromosome 14 of the C57 BL/6 mouse. A germ-line gene is considered to be functional if the coding region has an open reading frame (ORF) without a stop codon, and if there is no described defect in the splicing sites, RSS, and/or regulatory elements. A germ-line entity is considered to be an ORF if the coding region has an open reading frame but alterations have been described in the splicing sites, RSSs, and/ or regulatory elements, and/or changes of conserved amino acids have been suggested by the investigators to lead to incorrect folding and/or the entity is an orphon, a nonfunctional gene located outside the main chromosomal locus. A germ-line entity is considered to be a pseudogene if the coding region has a stop codon(s) and/or a frameshift mutation. In the case of V gene segments, these mutations may be either in the V gene coding sequence, or in the leader sequence. SOURCE: Gene numbers and definitions summarized from the International Immunogenetics Information System Web site: http://imgt.org. Accessed November 16, 2011. there are only three fully functional V␭ gene segments, although this number varies somewhat by strain (see Table 7-2). Interestingly, there are also three fully functional  chain constant regions, each one associated with its own J region segment (Figure 7-4b). The J-C4 segments are not expressed because of a splice site defect. Since recombination of Ig gene segments always occurs in the downstream direction (V to J), the location of the V1 variable region sequence upstream of the JC3 and JC1 segments but downstream from JC2 means that V1 is always expressed with either JC3 or JC1 but never with JC2. For the same reason, V2 is usually associated with the JC2 pairing, although occasional pairings of V2 with the 190 kb distant JC1 segments have been observed. In humans, 40% of light chains are of the  type and about 30 V-chain gene segments are used in mature antibody light chains (see Table 7-2). Downstream from the human V locus is a series of seven JC pairs, of which four pairs are fully functional. mouse strains). Just 0.7 kb downstream of the most 3 D segment is the JH region cluster, which contains four functional JH regions. A further gap separates the last JH segment from the first constant region exon, C1. The human VH locus has a similar arrangement with approximately 30 functional D segments, and 6 functional JH segments. Human D regions can be read in all three reading frames, whereas mouse D regions are mainly read in reading frame 1, because of the presence of stop codons in reading frames 2 and 3 in most mouse D regions. The eight constant regions of antibody heavy chains are encoded in a span of 200 kb of DNA downstream from the JH locus, as illustrated in Figure 7-4c. Recall that the constant regions of antibodies determine their heavy-chain class and, ultimately, their effector functions (see Chapter 13). Heavy-Chain Gene Organization Includes VH, D, JH, and CH Segments In V(D)J recombination, the DNA encoding a complete antibody V-region is assembled from V, D, and J (heavy chain) or V and J (light chain) segments that are initially separated by many kilobases of DNA. Each developing B cell generates a novel pair of variable region genes by recombination at the level of genomic DNA. Recombination is catalyzed by a set of enzymes, many of which are also involved in DNA repair functions (Table 7-3), and is directed to the appropriate sites on the Ig gene by recognition of specific DNA sequence motifs called Recombination Signal Sequences (RSSs). These sequences ensure that one of each Multiple VH gene segments lie across a region of approximately 3.0 Mb in both mice and humans (Figure 7-4c). These segments can be classified, like V segments, into families of homologous sequences. Humans express at least 38 functional VH segments and mice approximately 101 (see Table 7-2). Downstream from the cluster of mouse VH region segments is an 80 kb region containing approximately 14 D regions (the actual number varies among different The Mechanism of V(D)J Recombination The Organization and Expression of Lymphocyte Receptor Genes CHAPTER 7 233 type of segment (V and J for the light chain, or V, D, and J for the heavy chain) is included in the recombined variable region gene. During cleavage and ligation of the segments, the DNA is edited in various ways, adding further variability to the recombined gene. Proteins involved in V(D)J recombination TABLE 7-3 | Protein Function RAG-1/2 Lymphoid-specific complex of two proteins that catalyze DNA strand breakage and rejoin to form signal and coding joints TdT Lymphoid-specific protein that adds N region nucleotides to the joints between gene segments in the Ig heavy chain and at all joints between TCR gene segments HMG1/2 proteins Stabilize binding of RAG1/2 to Recombination Signal Sequences (RSSs), particularly to the 23-bp RSS; stabilize bend introduced into the 23-bp spacer DNA by the RAG1/2 proteins Ku70 and Ku80 heterodimers Binds DNA coding and signal ends and holds them in protein-DNA complex DNA PKcs In complex with Ku proteins, recruits and phosphorylates Artemis Artemis Opens the coding end hairpins XRCC4 Stabilizes and activates DNA ligase lV DNA ligase lV In complex with XRCC4, and Cernunnos ligates DNA ends Cernunnos With XRCC4, activates DNA ligase lV Recombination Is Directed by Signal Sequences If recombination is to occur in the DNA of every lymphocyte, then a mechanism must exist to ensure that it only occurs in antibody and T-cell receptor genes, and that the moving DNA segments end up exactly where they should be in the genome. Otherwise, dire consequences, including malignancy, can ensue. In the late 1970s, investigators working with light-chain genes, described two blocks of conserved sequences—a nonamer (a set of nine base pairs) and a heptamer (a set of seven base pairs)—that are highly conserved and occur in the noncoding regions upstream of each J segment. The heptamer appeared to end exactly at the J region coding sequence. Further sequencing showed that the same motif was repeated in an inverted manner on the downstream side of the V region coding sequences, again with the heptamer sequence ending flush with the V-region gene segment (Figure 7-5a). A further noteworthy feature of these sequences was the presence of a spacer sequence of either 12 or 23 bp between the heptamer and the nonamer. The significance of the spacer lengths was clear; they represented one, or two, turns (a) Nucleotide sequence of RSSs CACAGTG 23 bp ACAAAAACC GGTTTTTGT 12 bp CACTGTG GTGTCAC Heptamer 23 bp TGTTTTTGG Nonamer CCAAAAACA Nonamer 12 bp GTGACAC Heptamer 23 base pair spacer 12 base pair spacer FIGURE 7-5 Two conserved sequences (b) Location of RSSs in germ-line immunoglobulin DNA κ-chain DNA λ-chain DNA Heavy-chain DNA 5' L Vκ 3' 5' 12 5' 3' 23 L VH Cκ Jλ Cλ 23 L Vλ 3' Jκ 12 DH 23 12 12 JH 23 CH in light-chain and heavy-chain DNA function as recombination signal sequences (RSSs). (a) Both signal sequences consist of a conserved heptamer and conserved AT-rich nonamer; these are separated by nonconserved spacers of 12 or 23 bp. (b) The two types of RSS have characteristic locations within -chain, -chain, and heavychain germ-line DNA. During DNA rearrangement, gene segments adjacent to the 12-bp RSS can join only with segments adjacent to the 23-bp RSS. 234 PA R T I I I | Adaptive Immunity: Antigen Receptors and MHC (a) V cluster Germline DNA 5' 1 3' 3 1 Rearrangement between V3 and J3 5' 3' Recombined VJ gene in B cell 2 1 Constant region gene J cluster 2 3 3 2 3 4 Constant region gene 4 (b) D cluster (12-14 segments) V cluster Germline DNA J cluster (4 functional segments) Constant region gene 5' 3' D-J rearrangement 5' 3' Recombined VDJ gene in B cell 5' V-D rearrangement 3' FIGURE 7-6 Recombination between gene segments is required to generate a complete light chain gene. (a) Recombination between a V region (in this case, V3) and a J region (in this case J3) generates a single VC light-chain gene in each B cell. The recombinase enzymes recognize the RSS downstream of the V region (orange triangle) and upstream of the J region (brown triangle). In every case, an RSS with a 12-bp (one-turn) spacer is paired with an RSS with a 23-bp (two-turn) spacer. This ensures that there is no inadvertent V-V or J-J joining. (b) Recombination of V (blue), D (purple), and J (red) segments creates a complete heavy-chain variable region gene. Again, the recombinase enzyme recognizes the RSS sequences downstream of the V region, up- and downstream of the D region, and upstream of the J region, pairing 23-bp spacers with 12-bp spacers. of the double helix. Thus, the spacer sequence ensures that the ends of the nonamer and heptamer closest to the spacers would be on the same side of the double helix and, thus, accessible to binding by the same enzyme. The investigators correctly concluded that they had discovered the signal that directs recombination between the V and J gene segments and termed this heptamer 12/23 nonamer motif, the recombination signal sequence (RSS). To summarize, the RSS consists of three elements: sequence with a 12-bp spacer with one sequence with a 23-bp spacer, something we now know to be the case. Figures 7-6a and 7-6b illustrate the generation of complete light-chain and heavy-chain genes from individual V and J and V, D, and J segments, respectively. • Two proteins, encoded by closely linked genes, RAG1 (Recombination Activating Gene 1) and RAG2 (Recombination Activating Gene 2), were shown to be necessary for recombining antibody genes. The RAG1 and RAG2 genes are encoded just 8 kilobases apart and are transcribed together. The expression of RAG1 and RAG2 is developmentally regulated in both T and B cells (see Chapters 9 and 10) and, although RAG1 is expressed at all phases of the cell cycle, RAG2 is stable only in G0- or G1-phase cells. RAG1 is the predominant recombinase; it forms a complex with the RSS that is stabilized by the binding of RAG2. RAG2 by itself does not exhibit detectable RSS binding activity. • • An absolutely conserved, 7-bp (heptamer) consensus sequence 5-CACAGTG –3 A less conserved spacer of either 12 or 23 bp A second conserved, 9-bp (nonamer) consensus sequence 5-ACAAAAACC-3 In the heavy-chain gene segments, a similar pattern was noted. The spacer regions separating the heptamer and nonamer pairs were 23 bp in length following the V and preceding the J segments and 12 bp in length before and after the D segments. The relative locations of the 12 and 23 base pair spacers (Figure 7-5b) suggested that the VDJ recombinase enzyme is designed to bring together one Gene Segments Are Joined by the RAG1/2 Recombinase The Organization and Expression of Lymphocyte Receptor Genes V Binding of RAG1/2, HMG proteins Synapsis | CHAPTER 7 235 J 5' 3' 5' 3' 5' 3' RAG1/2 HMG proteins Cleavage and processing of signal and coding joints 5' 3' 5' 3' Artemis TdT DNA ligase IV NHEJ proteins Generation of functional Ig variable region gene 5' 3' Coding joint Signal joint FIGURE 7-7 Overview of recombination of immunoglobulin variable region genes. The RAG1/2 complex (represented together by the green oval) binds the RSSs and catalyzes recombination. Other enzymes fill in or cleave nucleotides at the coding end, and ligase completes the process. See text for details. [Adapted from Krangel, M. Nature Immunology 4, p. 625. 2003] Only three of the proteins implicated in V(D)J recombination are unique to lymphocytes: RAG1, RAG2, and Terminal deoxynucleotidyl Transferase (TdT), which is responsible for the generation of additional diversity in the CDR3 region of the antibody heavy chain (as we will see below). TdT is expressed only in developing lymphocytes and adds untemplated “N” nucleotides to the free 3 termini of coding ends following their cleavage by RAG1/2 recombinases. Other enzymes participating in the recombination process are not lymphoid specific. Whereas binding of the RSS by RAG1/2 can occur in the absence of any other proteins, other cellular factors, most of which are part of the NonHomologous End Joining (NHEJ) pathway of DNA repair are necessary for completion of V(D)J recombination. These other non-lymphocyte-specific proteins known to participate in V(D)J joining are described in Table 7-3. V(D)J Recombination Results in a Functional Ig Variable Region Gene The process of V(D)J recombination occurs in several phases (Figure 7-7). The end product of each successful rearrangement is an intact Ig gene, in which V and J (light chains) or V, D, and J (heavy chains) segments are brought together to create a complete heavy or light chain gene. The new joints in the antibody V region gene, created by this recombination process are referred to as coding joints. The joints between the heptamers from the RSSs are referred to as signal joints. The first phase of this process, DNA recognition and cleavage, is catalyzed by the RAG1/2 proteins. The second phase, end processing and joining, requires a more complex set of enzymatic activities in addition to RAG1/2, including Artemis, TdT, DNA ligase lV, and other NHEJ proteins. The individual steps involved in the process of recombination between V and J segments are shown sequentially in Figure 7-8. Step 1 Recognition of the heptamer-nonamer Recombination Signal Sequence (RSS) by the RAG1/RAG2 enzyme complex. The RAG1/2 recombinase forms a complex with the heptamer-nonamer RSSs contiguous with the two gene segments to be joined. Complex formation is initiated by recognition of the nonamer RSS sequences by RAG1 and the 12-23 rule is followed during this binding. Step 2 One-strand cleavage at the junction of the coding and signal sequences. The RAG1/2 proteins then perform one of their unique functions: the creation of a single-strand nick, 5 of the heptameric signal sequence on the coding strand of each V segment and a similar nick on the non-coding strand exactly at the heptamer-J region junction. (Figure 7-8 shows this process for the V segment only.) Step 3 Formation of V and J region hairpins and blunt signal ends. The free 3 hydroxyl group at the end of the coding strand of the VK segment now attacks the phosphate group on the opposite, non-coding VK strand, forming a new covalent bond across the double helix and yielding a DNA hairpin structure on the V-segment side of the break (coding end). Simultaneously, a blunt DNA end is formed at Step 1. RAG1/2 and HMG proteins bind to the RSS and catalyze synapse formation between a V and a J gene segment. V J 5' 3' 1 Step 5. Opening of the hairpin can result in a 5' overhang, a 3' overhang, or a blunt end. 5' 3' V 2 3 Artemis 5' V 5' 3' 3' 5' 3' Step 2. RAG1/2 performs a single stranded nick at the exact 5' border of the heptameric RSSs bordering both the V and the J segments. 5' Opening at 1 yields 5' overhang 5' 3' 5' 3' Coding end 5' 5' 3' Step 6. Cleavage of the hairpin generates sites for P nucleotide addition. V 3' 5' V 5' 3' V HO: P P AG AT J 3' 5' J TCGA 3' 5' ATAT P :OH V 5' 3' Signal end OH J 23 bp G T G A C A C 23 bp C A C T G T G P Nonamer P OH P OH Sequence at the signal junction results from the joining of the two heptameric regions Nonamer12 bp C A C T G T G C A C A G T G 23 bp Nonamer Nonamer12 bp G T G A C A C G T G T C A C 23 bp Nonamer TCGA TATA AGCT ATAT J 3' 5' Ligation of completed segments by DNA Ligase IV and NHEJ proteins 3' Nonamer 3' TA Filling in of complementary strands by DNA repair enzymes Step 4. Ligation of the signal ends P G T G T C A C 12 bp C A C A G T G 12 bp OH 3' 5' TC 3' 5' V 5' 3' Hairpin cleavage by Artemis 5' Step 7. Ligation of light chain V and J regions 5' Opening at 3 yields 3' overhang J 3' Step 3. The hydroxyl group attacks the phosphate group on the noncoding strand of the V segment to yield a covalently-sealed hairpin coding end and a blunt signal end. 3' Opening at 2 yields blunt end V TCGATATA AGCTATAT J 3' 5' FIGURE 7-8 The mechanism of V(D)J recombination. 1. The RAG1/2 complex (pale green oval) binds to the RSSs and catalyzes synapse formation. The coding (5 → 3) strand of DNA is drawn in a thick line, the non-coding (3 → 5) strand in a thin line. For steps 1 to 5 we show only the events associated with the V region gene segment, although the single-strand cleavage, hairpin formation, and templated nucleotide addition simultaneously occur at the borders of the V and the J segments. In this example, the V and J regions are encoded in the same direction on the chromosome, and so the DNA encoding the RSSs and the intervening DNA is released into the nucleus as a circular episome and will be lost on cell division. The DNA that was on the coding strand of the V region prior to rearrangement is emboldened. The actual signal joint is between the residues that were in contiguity with the V and J regions, respectively. Only the heptamer sequence is written out to preserve clarity. Nucleotides encoded in the germ-line genome are shown in black; P nucleotides are in blue, and non-templated nucleotides added by TdT at heavy chain VD and DJ joints are shown in red. Steps 8, 9, and 10, shown on the facing page, only occur in heavy chain loci. (See text for details.) The Organization and Expression of Lymphocyte Receptor Genes 5' D J TCGA 3' 3' 5' ATAT Step 8. In heavy chain VD and DJ joints only: Exonuclease cleavage results in loss of coding nucleotides at joint - can occur on either or both sides of joint 5' D J T 3' 3' 5' TAT Step 9. Non-templated nucleotides (in red) are added to the coding joint by TDT Step 10. Ligation of heavy chain by DNA ligase IV and NHEJ proteins 5' 3' D TCGTCTATA AGCAGATAT J 3' 5' the edge of the heptameric signal sequence. The same process occurs simultaneously on the J side of the incipient joint. At this stage, the RAG1/2 proteins and HMG proteins are still associated with the coding and signal ends of both the V and J segments in a postcleavage complex. Step 4 Ligation of the signal ends. DNA ligase IV then ligates the free blunt ends to form the signal joint. The involvement of particular enzymes in this process was deduced from observations of V(D)J recombination in natural and artificially generated systems lacking one or more enzymes (see Table 7-3). Step 5 Hairpin cleavage. The hairpins at the ends of the V and J regions are now opened in one of three ways. The identical bond that was formed by the reaction described in step 3 above, may be reopened to create a blunt end at the coding joint. Alternatively, the hairpin may be opened asymmetrically on the “top” or on the “bottom” strand, to yield a 5 or a 3 overhang, respectively. A 3 overhang is more common in in vivo experiments. Hairpin opening is catalyzed by Artemis, a member of the NHEJ pathway. Step 6 Overhang extension, leading to palindromic nucleotides. In Ig light-chain rearrangements, the resulting overhangs can act as substrates for extension DNA repair enzymes, leading to double stranded palindromic (P) nucleotides at the coding joint. For example, the top row of bases in the V region in the 5 to 3 direction reads TCGA. Reading backward on the bottom strand from the point of ligation also yields TCGA. The palindromic nature of the bases at this joint is a direct function of an asymmetric hairpin opening reaction. P nucleotide addition can also occur at both the VD and DJ joints of the heavy-chain gene segments, but, as described below, other processes can intervene to add further diversity at the VH-D and D-JH junctions. Step 7 Ligation of light-chain V and J Segments. Members of the NHEJ pathway repair both the signal and the coding joints, but the precise roles of each, and potentially other enzymes in this process, have yet to be fully characterized. | CHAPTER 7 237 During B-cell development, Ig heavy-chain genes are rearranged first, followed by the light-chain genes. This temporal dissociation of the two processes enables two additional diversifying mechanisms to act on heavy-chain V region segments. The enzymes responsible for these mechanisms are usually turned off before light-chain rearrangements begin. Comparative sequence analysis of germ-line and mature B-cell Ig genes demonstrated that both loss of templated nucleotides (nucleotides found in the germ-line DNA) and addition of untemplated nucleotides (nucleotides not found in germ-line DNA) could be identified in heavy-chain sequences. Two distinct enzyme-catalyzed activities are responsible for these findings. Step 8 Exonuclease trimming. An exonuclease activity, which has yet to be identified, trims back the edges of the V region DNA joints. Since the RAG proteins themselves can trim DNA near a 3 flap, it is possible that the RAG proteins may cut off some of the lost nucleotides. Alternatively, Artemis has also been shown to have exonuclease, as well as endonuclease activity, and could be the enzyme responsible for the V(D)J-associated exonuclease function. Exonuclease trimming does not necessarily occur in sets of three nucleotides, and so can lead to out-of-phase joining. V segment sequences in which trimming has caused the loss of the correct reading frame for the chain cannot encode antibody molecules, and such rearrangements are said to be unproductive. As mentioned above, such exonuclease trimming is more common at the two heavychain V gene joints (V-D and D-J) than at the light-chain V-J joint. In cases where trimming is extensive, it can lead to the loss of the entire D region as well as the elimination of any P nucleotides formed as a result of asymmetric hairpin cleavage. Step 9 N nucleotide addition. Non-templated (N) nucleotides are added by TdT to the coding joints of heavychain genes after hairpin cleavage. This enzyme can add up to 20 nucleotides to each side of the joint. The two ends are held together throughout this process by the enzyme complex, and again, loss of the correct phase may occur if nucleotides are not added in the correct multiples of three required to preserve the reading frame. Step 10 Ligation and repair of the heavy-chain gene. This occurs as for the light-chain genes. In describing V(D)J recombination, investigators must explain not only the mechanism of RSS recognition, cleavage, and ligation but also address the question of how two RSSs, located many kilobases distant from one another in the linear DNA sequence, are brought into close apposition. Furthermore, once successful recombination has occurred on one heavy-chain and one light-chain allele, this information must be communicated to the homologous chromosome, so that the other alleles can be silenced. How does this occur? Recent research indicates that both the structure and the location of recombinationally active, Ig V region DNA PA R T I I I 238 | Adaptive Immunity: Antigen Receptors and MHC change significantly as B-cell development proceeds, and that some of these changes are signaled by epigenetic alterations in the chromatin structure, mediated by specific methylation reactions on associated histone residues. DNA close to the nuclear membrane (pericentric DNA) does not include recombinationally active V regions; rather, DNA undergoing recombination moves away from the nuclear membrane, toward the center of the nucleus. The structure of the chromatin undergoing recombination also alters, such that long stretches of DNA are condensed into loops that allow the recombining sequences to come into closer contact with one another. Once successful recombination has occurred, the inactive allele has been shown to migrate to the pericentric regions. Thus, the recombination process is occurring within an actively modulating nucleoplasmic context, and is controlled by enzymes that alter chromatin structure, in addition to those that cleave and recombine the DNA. (a) V1 V2 V3 J1 J2 J3 J2 J3 J2 J3 J2 J3 V2 J2 J3 5' 3' Nicking and hairpin formation V1 V2 V3 J1 5' 3' Signal and coding joints form. Formation of signal joint results in circularizing of intervening DNA into an episome, that is lost upon subsequent cell divisions. V1 J2 J3 5' J1 V2 3' V3 (b) V1 V2 V3 J1 5' 3' Nicking and hairpin formation V1 V2 V3 J1 5' 3' Invert and ligate central fragment V1 5' Signal joint J1 V3 3' Note that the signal joint remains in the DNA upstream of the recombined antibody gene. FIGURE 7-9 Recombination can occur between DNA segments aligned in the same, or opposite, transcriptional direction on the chromosome. (a) Recombination is shown occurring between V1 and J2, which are both encoded in the same transcriptional orientation, from left to right. The intervening DNA is The direction of transcription is now consistent between the recombined V and J segments. excised as a circular episome. (b) Recombination is shown occurring between V2 and J2, which are encoded in opposite transcriptional orientations. In this case, the DNA containing the signal joint remains inverted in the DNA upstream of the recombined pair. The Organization and Expression of Lymphocyte Receptor Genes TABLE 7-4 | CHAPTER 7 239 Combinatorial antibody diversity in humans Nature of segment Number of heavy-chain segments (estimated) Number of ␬-chain segments (estimated) Number of ␭-chain segments (estimated) V 41 41 33 D 23 J 6 5 5 41  23  6  5658 41  5  205 30  5  165 Possible number of combinations Possible number of heavy-light chain combinations in the human  5658  (205 V(D)J Recombination Can Occur between Segments Transcribed in Either the Same or Opposite Directions Sequencing of chromosomal fragments containing multiple antibody V region genes showed that transcription of some V region genes occurs in a direction opposite from that of downstream D and J segments. However, work with artificial recombinase substrates has demonstrated that the same general mechanism of recombination is used, regardless of the transcriptional direction of the gene segments involved. When recombination occurs between two gene segments that are transcribed in the same direction, the intervening DNA is deleted and lost (Figure 7-9a). When they are transcribed in opposite directions, the DNA that was located between the V and J segments is retained, in an inverted orientation, in the DNA upstream of the rearranged VJ region (Figure 7-9b). Five Mechanisms Generate Antibody Diversity in Naïve B Cells This description of the complex and sophisticated apparatus by which Ig genes are created allows us to understand how such an immensely diversified antibody repertoire can be generated from a finite amount of genetic material. To summarize, the diversity of the naïve BCR repertoire is shaped by the following mechanisms: 1. 2. 3. 4. Multiple gene segments exist at heavy (V, D, and J) and light-chain (V and J) loci. These can be combined with one another to provide extensive combinatorial diversity (Table 7-4). P nucleotide addition results when the DNA hairpin at the coding joint is cleaved asymmetrically. Filling in the singlestranded DNA piece resulting from this asymmetric cleavage generates a short palindromic sequence. Exonuclease trimming sometimes occurs at the VDJ and VJ junctions, causing loss of nucleotides. Non-templated N nucleotide addition in heavy chains results from TdT activity. Mechanisms 2, 3, and 4 give rise to extra diversity at the junctions between gene segments, which contribute to CDR3. 5. 165)  2.09  106 In addition to these four mechanisms for generating antibody diversity that operate on individual heavy- or light-chain variable segments, the combination of different heavy and light-chain pairs to form a complete antibody molecule provides further opportunities for increasing the number of available antibody combining sites (Table 7-4). Combinatorial diversity: The same heavy chain can combine with different light chains, and vice versa. These five mechanisms are responsible for the creation of the diverse repertoire of BCRs and antibodies that is available to the organism before any contact with pathogens or antigen has occurred. Following antigenic stimulation, B cells are able to use yet another mechanism, unique to the immune system, to further diversify and refine the antigenspecific receptors and antibodies: somatic hypermutation. As described in Chapter 12, a specialized enzyme complex targets the genes encoding the variable regions of Ig genes only in those B cells that have undergone antigen-specific activation in the presence of T-cell help. B cells are exposed to successive cycles of mutation at the BCR loci, followed by antigen-mediated selection in specialized regions of the lymph node and spleen. The end result of this process is that the average affinity of antigen-specific BCRs and antibodies formed at the end of an immune response is considerably higher than that at its instigation. This process is referred to as affinity maturation. This section has described the process of the generation of the primary Ig variable region repertoire as it occurs in humans and rodents. Although the same principles apply to most vertebrate species, different species have evolved their own variations. For example, the process of gene conversion is used in chickens, and some species, such as sheep and cows, use somatic hypermutation in the generation of the primary as well as the antigen-experienced repertoire. In Evolution Box 7-2, we describe the evolution of this system of recombined lymphocyte receptors, addressing the current hypothesis that the key event was the introduction of the RAG1/2 gene segment into the early vertebrate genome as a transposon. 240 PA R T I I I | Adaptive Immunity: Antigen Receptors and MHC EVOLUTION Evolution of Recombined Lymphocyte Receptors Scientists are slowly beginning to • • • The DNA-binding region of RAG1 is strikingly homologous to that of known transposable elements and is evolutionarily related to members of the Transib transposase family, currently expressed in species such as fruit flies, mosquitoes, silkworm, and the red flour beetle. The inverted repeats in the RSSs of the Ig and TCR gene segments are structurally similar to those found in other transposons. The mechanism of action of the RAG proteins involves formation of a DNA hairpin intermediate reminiscent of the action of certain transposases. For example, another phylogenetically con- • Current thinking supports the hypothesis that the appearance of the transposon in a primordial antigen receptor gene might have separated the receptor gene into two or more pieces. This hypothesis is supported by the presence, in lower chordates, of genes that are related to BCR and TCR V regions, and that therefore could RAG1 and RAG2 VLR TCR, BCR, and MHC Adaptive immunity based on rearranging antigen receptors TLR, NLR, and SR Million years ago Mammals (humans) Birds (chickens) Reptiles (snakes) Amphibians (frogs) Bony fish (sturgeons) Bony fish (zebrafish, medaka fish) Cartilaginous fish (sharks) Cyclostomes (lampreys) Cyclostomes (hagfish) Urochordates (sea squirts) Cephalochordates (amphioxi) Hemichordates (acorn worms) Innate immunity based on pattern-recognition receptors Echinoderms (sea urchins) understand how the process of V(D)J recombination may have evolved. With that understanding has come an appreciation for the fact that the BCRs and TCRs of the vertebrate adaptive immune system may not be the only immune receptor molecules generated by recombinatorial genetics. Ig-like receptors have been identified in species as ancient as the earliest jawed vertebrates. However, extensive analyses of the only surviving jawless vertebrates (Agnathans), the hagfish and lamprey, show no evidence of TCR or BCR V(D)J segments, RAG1/2 genes, or genes encoding elements of a primitive Major HistoCompatibility Complex (MHC) system. This suggests that the RSS-based, recombinatorially generated set of adaptive immune receptors first emerged in a common ancestor of jawed vertebrates, most probably more than 500 million years ago (Figure 1). The driving force in the development of the adaptive immune system was most probably the incorporation of an ancestor of the RAG1 genes into the ancestral genome in the form of a transposable element. This hypothesis is supported by several observations and experiments: served transposon in flies, the HERMES transposon, has been shown to induce a double-strand break via the hairpinformation mechanism described above for V(D)J recombination. RAG proteins have the ability to transpose an RSS-containing segment of DNA to an unrelated target DNA in vitro. 222 326 370 476 525 652 794 891 896 FIGURE 1 The emergence of the RSS-dependent recombination-based adaptive immune system coincides with the first appearance of the jawed vertebrates. SR  scavenger receptor. [Adapted from Figure 1, M. F. Flajnik and M. Kasahara, 2010, Origin and evolution of the adaptive immune system: Genetic events and selective pressures, Nature Reviews Genetics 11:47–59.] | The Organization and Expression of Lymphocyte Receptor Genes CHAPTER 7 241 BOX 7-2 have provided the substrate for the transposon invasion. When the transposon subsequently jumped out of the gene, the RSSs would have been left behind. Multiple such rounds of transposition followed by gene duplication would have given rise to the TCR and BCR loci we know today. Interestingly, our knowledge of phylogeny suggests that two waves of gene duplications occurred at the time of vertebrate origin, right around the time of the proposed transposon entry. Much still remains unknown about the evolution of the adaptive immune receptor molecules. What was the nature of the primordial antigen receptor that was the target for the first transposition? At what stage did the RAG2 gene become associated with the RAG1 component? And how do we account for the evolution of the Ig receptors of species such as the shark, in which fully rearranged V(D)J genes exist in the germ-line genome? Recent evidence has suggested that the recombinatorial strategy for generating antigen receptors is not limited to the RSS-based Ig and TCR systems of vertebrates. Jawless fish, such as lampreys and hagfish, possess lymphocyte-like cells that can be stimulated to divide and differentiate. Moreover, these cells release specific agglutinins after immunization with antigens, and higher serum levels of these agglutinins are secreted after a secondary, than after a primary, immunization, suggesting that immunological memory exists in these fish. These lymphocyte-like cells therefore appear to express all the hallmarks of adaptive immunity, and yet the agglutinins they secrete do not appear to have an Ig-like structure. Analysis of the agglutinins has revealed that activated lamprey “lymphocytes” express abundant quantities of leucinerich-repeat (LRR)-containing proteins. LRRs are protein motifs that are frequently associated with protein-protein recognition, and such motifs have already been encountered in the context of the Toll-like receptors described in Chapter 5. The lamprey agglutinin receptors are gener- (a) Amino acids/cassette SP LRRNT LRR1 27-34 24 LRRV (0-8) GPI Stalk LRRVe CP LRRCT 24 24 HP 16 48-63 (b) 3 2 4 1 5' LRR cassettes 5 3' LRR cassettes Functional VLR LRRNT LRR1 LRRV1 LRRV2 LRRV3 LRRVe LRRCP LRRCT ated by recombination of gene segments during lymphocyte development, and the LRR receptor repertoires of these fish are remarkably diverse. These receptors have therefore been named variable lymphocyte receptor (VLR) molecules. Figure 2a shows an example of the arrangement of the protein modules in one of these VLR molecules. At the amino terminal of the protein is an invariant signal peptide, followed by a 27–34 residue N-Terminal Leucine-Rich Repeat (LRRNT). This is followed by the first of several FIGURE 2 Agnathans such as lampreys and hagfish have evolved a recombination-based system of recognition molecules that does not depend on RSSs. Agnathans, such as lamprey and hagfish, have lymphocyte-like cells which carry antigen receptors that are remarkably diverse and are created by a system of recombination of small gene cassettes. (a) The modules that make up a complete VLR receptor. (See text for details.) (b) The assembly of a complete VLR gene from the basic skeleton germ-line and flanking cassettes. (c) The three-dimensional structure of a VLR, showing the repeating arrangement of the randomly recombined LRR protein modules. [Adapted from Figures 1 and 2, Herrin, B. R., and M. D. Cooper. 2010. Journal of Immunology 185:1367.] 24-residue Leucine-Rich Repeat (LRR) modules, LRR1, connected to a series of up to eight 24-residue LRR modules with variable sequences (LRRVs). At the C-terminal of the molecule, a 24-residue end LRRV segment (LRRVe) is attached to a short 16-residue connecting peptide (CP), which culminates in a 48–63 residue C-terminal LRR (LRRCT). The VLR molecules are attached to the lymphocyte membrane by an invariant stalk, rich in threonine and proline residues, connecting to a glycosyl-phosphatidyl-inositol (continued) 242 PA R T I I I EVOLUTION | Adaptive Immunity: Antigen Receptors and MHC BOX 7-2 (continued) (GPI) anchor and a hydrophobic peptide. Lymphocyte activation leads to phospholipase cleavage at the GPI anchor, enabling soluble forms of the VLRs to be released from the lymphocyte following antigen stimulation. The assembly process that generates the completed protein occurs only in lymphocytes. In the germ-line DNA, LRR cassettes flank the skeleton LRR genes, which consist initially only of sequences coding for parts of the LRRNT and LRRCT, separated by non-coding DNA (Figure 2b). During lymphocyte development, the non-coding sequence is replaced with variable LRRs. Gene segments are copied from one part of the genome into another in a one-way process, similar to gene conversion. Each lamprey lymphocyte expresses a unique VLR gene from a single allele, and the diversity of the repertoire is limited only by the number of lymphocytes. Figure 2c shows a cartoon of the structure of one of these receptors, generated from an x-ray crystallographic model. Comparison of the sequences of several of these primarily -stranded structures has shown that their sequence diversity is concentrated on the concave (left) side of the molecule and can be attributed to the inherent diversity of the LRR cassettes. Did the two LRR- and V(D)J-based, recombinatorially derived sets of immune receptors exist side by side in an ancestral species? This we do not know. Agnathans other than lampreys and hagfish were extinguished 400 million years ago, and B-Cell Receptor Expression The expression of a receptor on the surface of a B cell is the end result of a complex and tightly regulated series of events. First, the cell must ensure that the various gene recombination events culminate in productive rearrangements of both the heavy- and light-chain loci. Second, only one heavy-chain and one light-chain allele must be expressed in each B cell. Finally, the receptor must be tested to ensure it does not bind self antigens, in order to protect the host against the generation of an autoimmune response. Pre-BCR J reB Vp λ5 The random nature of Ig heavy- and light-chain gene rearrangement means that more than one heavy-light-chain pair could potentially be expressed on the surface of individual B cells. Furthermore, since each heavy chain could potentially combine with both light chains and vice versa, this could result in the creation of B cells bearing a variety of different antigen-binding sites. Whereas this opportunity to increase the number of available receptors may initially sound advantageous to the organism, in practice the presence of more than one receptor per B cell creates prohibitive difficulties for those mechanisms that protect against the generation of autoimmunity. The mechanism by which B cells ensure that only one heavy- and one lightchain allele are transcribed and translated is referred to as allelic exclusion. The rearrangement of Ig genes occurs in an ordered way, and begins with recombination at one of the two homologous chromosomes carrying the heavy-chain loci. The production of a complete heavy chain and its expression on the B-cell surface in concert with a surrogate light chain, made up of two proteins, VpreB and 5 (Figure 7-10), signals the end of heavy-chain gene rearrangement and, thus, only one antibody heavy chain is allowed to complete the rearrangement process. The mechanism by which this allelic exclusion occurs is still under investigation. However, we do know that following successful arrangement at one allele, the VH gene locus on the other chromosomes is methylated and recruited to heterochromatin. VD Allelic Exclusion Ensures That Each B Cell Synthesizes Only One Heavy Chain and One Light Chain we have access only to the fossil remains of ostracoderms (agnathans with dermal skeletons), which are thought to be the ancestors of the gnathostomes. One theory would hold that the VLR recombinatorial system evolved in an ancestor common only to hagfish and lampreys, and could have been causal in their ability to survive when other agnathan species succumbed to environmental insult. Alternatively, the VLR and Ig-based systems may have co-existed for a time, with lymphocytes bearing both types of receptors, until one of the systems was lost. Combinatorial genetics in the generation of the vertebrate antigen receptors may not be as unique a mechanism as immunologists had previously thought. FIGURE 7-10 Ig heavy-chain expression on the cell surface in concert with VpreB and ␭5 signals, via Ig␣/Ig␤, the end of heavy-chain rearrangement. | The Organization and Expression of Lymphocyte Receptor Genes Pre-BCR, made up of μ heavy chain plus the surrogate light chain, inhibits rearrangement of μ allele #2 and induces κ rearrangement CHAPTER 7 243 μ + κ chains inhibit rearrangement of κ allele #2 and λ rearrangement IgM μ + κ chains inhibit λ rearrangement Pre-BCR μ+κ μ VH DH JH μ + λ chains inhibit rearrangement of λ allele #2 JH Productive allele #1 Vκ Jκ Productive allele #1 μ+κ Vκ Jκ μ μ Productive allele #2 Nonproductive allele #1 VH DH Progenitor B cell DH Productive allele #2 μ+λ JH JH DH VH VH JH DH Nonproductive allele #1 μ Vλ Jλ Nonproductive allele #2 Productive allele #1 μ Vλ Jλ Nonproductive allele #1 μ+λ Productive allele #2 μ Nonproductive allele #2 Nonproductive allele #2 Cell death Cell death FIGURE 7-11 The generation of a functional immunoglobulin receptor requires productive rearrangement at heavy- and light-chain alleles. Heavy-chain genes rearrange first, and once a productive heavy-chain gene rearrangement occurs, the pre-BCR containing the  protein product prevents rearrangement of the other heavy-chain allele and initiates light-chain rearrangement. In the mouse, rearrangement of  light-chain genes precedes rear- If the first attempt at a heavy-chain rearrangement is unproductive (i.e., results in the formation of out-of-frame joints caused by P and N nucleotide addition or by exonuclease trimming at the joint), rearrangement will initiate at the second heavy-chain allele. If this is also unsuccessful, the B cell will die. Once a complete heavy chain is expressed, lightchain rearrangement begins. In mice, light-chain rearrangement always begins on one of the  alleles, and continues until a productive light-chain rearrangement is completed. In humans, it may begin at either a  or a  light-chain locus. Because successive arrangements (using upstream V regions and downstream J regions) can occur on the same light-chain chromosome (see below) and there are four alleles (two  and two ) to choose from, once a B cell has made a complete heavy chain, it will normally progress to maturity. Successful completion of light-chain rearrangement results in the expression of the BCR on the cell surface, and this signals the end of further Ig receptor gene rearrangement events (Figure 7-11). rangement of the  genes, as shown here. In humans, either  or  rearrangement can proceed once a productive heavy-chain rearrangement has occurred. Formation of IgM inhibits further light-chain gene rearrangement. If a nonproductive rearrangement occurs for one allele, then the cell attempts rearrangement of the other allele. [Adapted from G. D. Yancopoulos and F. W. Alt, 1986, Regulation of the assembly and expression of variable-region genes, Annual Review of Immunology 4:339.] The process of generating a fully functional set of B cells is energetically expensive for the organism because of the amount of waste involved. On average, two out of three attempts at the first heavy-chain chromosomal locus will result in an unproductively rearranged heavy chain, and the same is true for the rearrangement process at the second heavy-chain locus. Therefore the probability of successfully generating a functional heavy chain is only 1/3 (at the first allele) {2/3 (probability the first rearrangement was not successful)  1/3 (probability that the second rearrangement is successful)}  0.55 or 55%. Receptor Editing of Potentially Autoreactive Receptors Occurs in Light Chains The immune system can still make some changes even after the completed receptor is on the cell surface. If the receptor is found to be auto-reactive, the cell can swap it out for 244 PA R T I I I Germline DNA | 5' 3' Adaptive Immunity: Antigen Receptors and MHC V cluster 1 2 Constant region J cluster 1 3 2 3 4 First rearrangement between V3 and J3 is productive, but combination of the resultant light chain with Ab heavy chain results in an autoimmune antibody Recombined VJ gene in B cell 5' 1 3' 3 3 2 4 Second rearrangement between V2 and J4 is productive, and combination with the Ab heavy chain results in an non autoimmune antibody 5' 3' 1 2 4 FIGURE 7-12  light-chain receptor editing. In the event of a productive light-chain rearrangement that leads to the formation of an auto-antibody, additional rounds of rearrangement may occur between upstream V and downstream J gene segments. Here, the primary rearrangement between V3 and J3 is edited out with a secondary rearrangement between V2 and J4. a different one, in a process termed receptor editing. Since the receptor’s binding site contains elements of both the heavy and the light chain, changing just one of these is often sufficient to alter the specificity, and light-chain receptor editing has been shown to occur quite frequently. In this process, the DNA rearrangement machinery is switched back on after the first completed receptor in the immature B cell has been found to have autoimmune specificity. For example, as shown in Figure 7-12, the initial rearrangement of V3 to J3 may result in an Ig receptor molecule that bound to a self antigen. The B cell can then rearrange a light-chain gene at the second -chain allele, it can rearrange a -chain allele, or it can edit the original -chain rearrangement. All of these choices have been demonstrated in different B cells. However, the term receptor editing refers to the process in which a B cell uses the same allele more than once, and engages in a second rearrangement process in which a variable region gene segment upstream of the original segment (in this case V2) is recombined with a J segment downstream of the initially utilized J region (in this case, J4). Switching out the rearranged VH segments is less common. Since rearrangement of the heavy-chain variable region deletes the RSSs on either side of the D region sequences, it was at first thought to be impossible. However, recent advances have demonstrated that cryptic RSSs exist that can be used in heavy-chain editing, and so some measure of heavy-chain receptor editing does occur. Ig Gene Transcription Is Tightly Regulated Immunoglobulin genes are expressed only in cells of the B lymphocyte lineage, and within this lineage, expression of Ig genes is developmentally regulated (see Chapter 10). The control of Ig gene transcription is therefore necessarily complex, although it involves the familiar paradigm of transcription factors that bind to sequence elements in the DNA (Table 7-5 and Figure 7-13). Expression of the Ig genes requires the coordination of two types of cis regulatory elements: promoters and enhancers. Like other promoters, the Ig heavy and lightchain gene promoters contain a highly conserved, AT-rich sequence called the TATA box, which serves as the binding site for RNA polymerase. RNA polymerase II starts transcribing the DNA from the initiation site, about 25 base pairs (bp) downstream of the TATA box. Each VH and VL gene segment has a promoter located just upstream from its leader sequence. The promoters upstream of Ig VH and VL gene segments bind to RNA polymerase II quite weakly, and the rate of transcription of VH and VL coding regions is very low in unrearranged germ-line DNA. Rearrangement of Ig genes brings the promoters close enough to enhancers located in the intron sequences and in regions downstream of the C region genes for them to influence the V region promoters. Consequently, the rate of transcription is increased by a factor of 104 upon V gene rearrangement. However, although the transcription of unrearranged genes is low, it is not absent in developing B cells; evidence now suggests that, as the Ig genes move out of the pericentric regions of the nucleus and assume more euchromatic configurations prior to recombination, a low level of transcription occurs which indicates to the recombination machinery that they are in a state of readiness for rearrangement to occur. | The Organization and Expression of Lymphocyte Receptor Genes TABLE 7-5 CHAPTER 7 245 Some of the cis regulatory elements that control immunoglobulin gene transcription Elements Function Location(s) in Ig genes Promoters Promote initiation of transcription of neighboring gene in a specific direction. Bind RNA polymerase and direct the formation of the pre-initiation complex. Upstream of VH, D, V, V1, and V2. Stimulate transcription of associated genes/gene segments when bound by transcription factors (TFs). Can function at variable distance from promoter, and in either orientation. Intronic enhancers lie between JH and C and between J and C. Enhancers Symbol in Figure 7-13 Additional promoters are found upstream of all constant regions. Arrow denotes direction of transcription. 3 enhancers lie to the 3 end of the C gene and may be part of a locus control region. Light-chain enhancers lie 3 of C and C genes. Locus control regions Collection of smaller elements that may each have enhancer function. Complex regulatory region, RR, 3 of murine C gene. Contains four enhancers in two separate units. The J cluster and each of the D genes of the heavy-chain locus are also preceded by promoters. These D and J promoters allow for transient transcription of DJ and J segments, which appears to be important for efficient VDJ rearrangement. So far, there is no evidence of conventional promoter sequences upstream of any of the J loci, although some germ-line transcription of J sequences has been detected. The control of Ig gene transcription depends on the interplay between a large number of transcription factors that bind to the promoter and enhancer regions of the Ig genes, and binding of different combinations of these proteins to VH promoters IgH 5' 3' Constant region promoters μ Jκ regions κ intron enhancer γ1 3' enhancer/LCR/RR α 3' κ enhancer 5' 3' Vλ2 promoter Igλ μ intron enhancer D region promoters Vκ promoters Igκ promoters and enhancers can lead to very different outcomes. In Chapter 10, we discuss how transcription of Ig genes is controlled by varying combinations of transcription factors during B cell development. Occasionally, the powerful B-cell Ig enhancers can engage in activities that are less than benign. Translocation of the c-myc oncogene close to the Ig enhancer regions results in constitutive expression of the c-Myc protein and the generation of an aggressive, highly proliferative B-cell lymphoma. Similarly, the translocation of the bcl-2 gene close to the Ig enhancer leads to suspension of programmed cell death in B cells, resulting in follicular B-cell lymphoma. 5' 3' Vλ2 JCλ2 λ2-4 JCλ4 enhancer Vλ1 promoter Vλ1 JCλ3 λ3-1 enhancer JCλ1 FIGURE 7-13 Locations of the cis regulatory control elements on immunoglobulin genes. The expression of Ig genes is controlled by transcription factor binding to promoter and enhancer regions. Promoters are designated by rectangular arrows and enhancer regions by green ovals. Accessibility to recombination is often preceded by low levels of transcription from downstream D and J segments. 246 PA R T I I I | Adaptive Immunity: Antigen Receptors and MHC Mature B Cells Express Both IgM and IgD Antibodies by a Process That Involves mRNA Splicing As described above, the same Ig heavy-chain variable region can be expressed in association with more than one heavychain constant region. Furthermore, each of the heavy-chain constant regions can be synthesized as both membranebound and secreted forms. Immature B cells in the bone marrow express only membrane IgM receptors. However, as the B cells mature, they also place IgD receptors on their cell surface. These IgD receptors bear the identical antigen-binding V region, but carry a , rather than a  constant region. In this section, we describe how cells create the membrane-bound and secreted forms of IgM and IgD by RNA splicing. The generation of antibodies belonging to the heavy-chain classes other than IgM and IgD occurs by an additional DNA recombination process, referred to as class switch recombination (CSR), and results in the loss of the DNA encoding the  and gene segments. CSR is described in Chapter 12, because it is a process that occurs only after antigen-mediated activation of B cells. The primary heavy-chain transcript in a B cell that has not yet been exposed to antigen encodes both the C and the C constant regions, and is approximately 15 kb in length. The determination of whether a cell will synthesize the membrane-bound or secreted forms of  and/or heavy chains is made by nuclear proteins that select particular choices of polyadenylation and RNA splicing sites. During mRNA processing, polyadenylation sites signal for the cleavage and addition of a poly-A sequence at a particular site on the primary transcript, that then becomes the 3 terminus of the mature, processed mRNA. (Poly-A tails are characteristically found at the 3 termini of many eukaryotic mRNA species.) A nuclear complex called the spliceosome determines the sites at which the primary transcript will be cleaved and spliced to form the mature mRNA. Figure 7-14 illustrates the splicing and polyadenylation choices made when a cell generates membrane bound IgM (the only Ig species expressed by immature B cells) versus membrane-bound IgD, which is expressed upon B-cell maturation. Both of these species are present on the cell membrane prior to antigen encounter. Analysis of the primary Ig transcript revealed that the C4 exon (the 3-most exon of the  heavy chain) ends in a nucleotide sequence called S (secreted), which encodes the hydrophilic portion of the CH4 domain of secreted IgM. The S portion of the mRNA sequence includes a polyadenylation site, polyadenylation site 1. Two additional exons, M1 and M2 (M for membrane), are located 1.8 kb downstream of the 3 end of the C4 exon. M1 encodes the membranespanning region of the heavy chain. M2 encodes the three cytoplasmic amino acids at the C-terminus of IgM, followed by polyadenylation site 2 (see Figure 7-14a). Production of mRNA encoding the membrane-bound form of the  chain occurs when cleavage of the primary transcript and addition of the poly-A tail occurs at polyadenylation site 2 (see Figure 7-14b). RNA splicing then removes the S sequence at the 3 end of the C4 exon, and joins the remainder of the C4 exon to the M1 and M2 exons. The RNA sequence between the M1 and M2 exons is also spliced out, resulting in the production of mRNA encoding the membrane-bound form of the IgM heavy chain. This is the only splicing pattern that occurs in immature B cells. As the B cell matures, it begins to express membranebound IgD in addition to membrane IgM (see Figure 7-14c). The exons encoding the membrane-bound and secreted forms of IgD are arranged similarly to those of IgM, with polyadenylation sites 3 and 4 at the 3 termini of the sequences encoding the secreted and membrane-bound forms of IgD respectively. If polyadenylation occurs at site 4, and the C RNA is spliced out, the cell will make the membrane-bound chain, and express IgD as well as IgM on the cell surface. Since mature B cells bear both IgM and IgD on their surfaces, it is clear that both splicing patterns can occur simultaneously. Following antigen stimulation, the B cell begins to generate the secreted form of IgM (Figure 7-15a). Polyadenylation shifts to site 1, and the mRNA sequences downstream of site 1 are degraded. (Similarly, the use of splice site 3 generates secreted IgD, although this is not shown for purposes of clarity.) The secreted form of IgD is used less frequently than the secreted form of the other antibody heavy-chain classes, and so splice site 3 is used relatively rarely. At the protein level, the carboxyl-terminus of the membrane-bound form of the Ig  chain consists of a sequence of about 40 amino acids, which is composed of a hydrophilic segment that extends outside the cell, a hydrophobic transmembrane segment, and a very short hydrophilic segment at the carboxyl end that extends into the cytoplasm (Figure 7-15b). In the secreted form, these 40 amino acids are replaced by a hydrophilic sequence of about 20 amino acids in the carboxyl terminal domain. What mechanism determines which splice sites are used? This question is not yet fully resolved. However, work with artificial constructs containing different combinations of the various splice sites suggests that there is an intrinsic “strength” to each of the splice sites that determines how frequently it is used, and that this strength reflects the binding affinity between the DNA sequence at the splice site and the spliceosome. Analysis of the sequences of the splice sites in Ig mRNA suggests that splice site 2 is considerably “stronger” than splice site 1, which explains why immature B cells make a preponderance of the membrane form of the  chain. B-cell stimulation appears to cause an increase in the concentration of the splice site recognition proteins, as well as the early termination of transcription prior to the M1 and M2 sequences, thereby facilitating the use of splice site 1 and | The Organization and Expression of Lymphocyte Receptor Genes CHAPTER 7 247 (a) H-chain primary transcript Cδ Cµ L VDJ µ1 J µ2 µ3 µ4 S δ1 M1 M2 δ2 δ3 S M1M2 5' 3' - 6.5 kb Poly-A site 1 Poly-A site 3 Poly-A site 2 Poly-A site 4 (b) Polyadenylation of primary transcript at site 2 → µm Cµ L VDJ J S M1 M2 µm transcript 5' (A)n Splicing L µm mRNA VDJ µ1 µ2 µ3 µ4 M1 M2 (A)n 5' (c) Polyadenylation of primary transcript at site 4 → δm Cµ VDJ L J µ1 µ2 µ3 µ4 S δm transcript 5' Cδ M1 M2 δ1 δ2 δ3 S M1M2 (A)n Splicing L δm mRNA 5' VDJ δ1 δ2 δ3 M1 M2 (A)n FIGURE 7-14 Differential expression of membrane forms of ␮ and ␦ heavy chains by alternative RNA processing. (a) Structure of rearranged heavy-chain gene showing C and C exons and poly-adenylation sites. (b) Structure of m primary transcript and m mRNA resulting from polyadenylation at site 2, followed by RNA splicing. (c) Structure of m primary transcript and m mRNA resulting from polyadenylation at site 4 and RNA splicing. Both processing pathways can proceed in any given B cell. generation of the secreted form of the protein. A similar mechanism may control the differential splicing of the membrane-bound and secreted forms of the chain. T-Cell Receptor Genes and Expression The initial characterization of the BCR molecule was facilitated by the fact that the secreted Ig product of the B cell shares the antigen-binding region with the membrane receptor. The secreted Ig proteins could be used as antigens to generate antibodies in other species that recognized the B-cell surface receptor and these anti-receptor antibodies could then be employed as reagents to isolate and characterize the BCR. In addition, a set of monoclonal B-cell tumors secreting high concentrations of soluble antibodies was available. Investigators attempting to purify and analyze the T-cell receptor (TCR) enjoyed no such advantages, and therefore it should not be surprising to learn that several decades elapsed between the characterization of the BCR and that of its T-cell counterpart. Indeed, characterization of the TCR was so difficult and vexing that one of the distinguished writers for the scientific journal Nature, Miranda Robertson, referred to it as “The Hunting of the Snark,” in reference to the Lewis Carroll poem of the same name, which described the search for a mythical creature. In Chapter 3, we described the TCR protein. Below, we describe the parallel story of the discovery of the TCR genes. Understanding the Protein Structure of the TCR Was Critical to the Process of Discovering the Genes In Chapter 3 we described how investigators were able to use a monoclonal antibody (mAb) specific for the TCR to purify and characterize it as an heterodimeric protein. Each chain was shown to consist of one variable and one constant region domain, and both chains contained regions arranged 248 PA R T I I I | Adaptive Immunity: Antigen Receptors and MHC (a) Cµ L Primary H-chain 5' transcript VDJ J µ1 µ2 µ3 Cδ µ4 S M1 M2 3' Poly-A site 1 Poly-A site 2 Poly-A site 3 Poly-A site 4 Polyadenylation Site 1 RNA transcript for secreted µ L V D J J µ1 µ2 µ3 Site 2 RNA transcript for membrane µ µ1 µ2 µ3 µ4 S M1 M2 L V DJ J µ4 S (A)n 5' (A)n 5' RNA splicing L V D J µ1 µ2 µ3 µ4 M1 M2 L V D J µ1 µ2 µ3 µ4 S 5' 5' (A)n mRNA encoding secreted µ chain (A)n mRNA encoding membrane µ chain Cμ4 (b) CHO SS bridge T G K P T L Y N V S L I M S D T G G T C Y 556 Key: Cμ4 Cμ 4 Hydrophilic + Hydrophobic 556 556 563 Encoded by S exon of Cμ 576 – COOH Encoded by M1 and M2 exons of Cμ 575 576 COOH Outside 568 576 Membrane 594 597 Cytoplasm COOH T 556 – E G – E V – A N E – E – E G F 568 – E N L F I W V T T S A T L F S T T L Y LV F L S T L F 594 + K V + K 597 COOH Secreted μ FIGURE 7-15 Differential expression of the secreted and membrane-bound forms of immunoglobulin ␮ chains is regulated by alternative RNA processing. (a) Structure of the primary transcript of a rearranged heavy chain, showing the C exons and poly-A sites. Polyadenylation of the primary transcript at either site 1 or site 2, and subsequent splicing (indicated by V-shaped lines) generates mRNAs encoding either secreted or Membrane μ membrane  chains, as shown. (b) Amino acid sequences of the carboxyl terminal ends of secreted and membrane  heavy chains. Residues are described by the single-letter amino acid code. Hydrophilic (pink) and hydrophobic (yellow) residues and regions are shown, charged amino acids are indicated with a or , and the N-linked glycosylation site on the secreted form is labeled “CHO.” The Organization and Expression of Lymphocyte Receptor Genes • • Vα Cα Vβ Cβ • • [PDB ID 1TCR.] in the characteristic Ig fold (Figure 7-16). The presence of these variable and constant regions provided clues as to the potential arrangement of the genes that encode the TCR proteins, and enabled researchers to design the experiments that ultimately led to the characterization of the TCR genes. The -Chain Gene Was Discovered Simultaneously in Two Different Laboratories Two different research groups published papers in the same issue of Nature in 1984 that described the discovery of the TCR genes in mice and in humans respectively. Mark Davis, Stephen Hedrick, and their collaborators utilized a uniquely creative approach to isolate the receptor genes from several mouse T-cell hybridoma cell lines, which we describe below. The other group, headed by Tak Mak, used more classical methods of genetic analysis to isolate the TCR genes from human cell lines. The Davis-Hedrick group strategy was based on four assumptions about TCR genes: • • TCR genes will be expressed in T cells but not B cells. Since the genes encode a membrane-bound receptor, the transcribed mRNA will be found associated with membrane-bound polyribosomes. CHAPTER 7 249 Like Ig genes, TCR genes will code for a variable and a constant region. Like Ig genes, the genes that encode the TCR will undergo rearrangement in T cells. Prior experiments had revealed that only about 2% of the genes expressed by lymphocytes differed between T and B cells. Furthermore, only a small proportion (about 3%) of T-cell mRNA had been shown to be associated with membrane-bound polyribosomes. Davis and Hedrick reasoned that if they could generate 32P-labeled DNA copies (cDNA) of the membrane-bound polyribosomal fraction of antigen-specific T-cell mRNA, and remove from that population those sequences that were also expressed in B cells, they would be left with labeled, T-cell-derived cDNA that would be greatly enriched in sequences encoding the TCR (Figure 7-17). They therefore performed the following steps: • FIGURE 7-16 The heterodimeric ␣␤ T-cell receptor. The alpha chain is shown here in green and the beta chain in blue. | • • Extract membrane-bound polysomal mRNA (RNA sequences that encode membrane-bound proteins) from several T hybridoma lines. Reverse transcribe the polysomal mRNA in the presence of 32P-labeled nucleotides, to generate radioactive cDNA copies of each of the mRNA species. Mix the 32P-labeled T-cell cDNA with B-cell mRNA and remove all the cDNA that hybridized with B-cell mRNA. (This leaves behind only the T-cell mRNA that encodes proteins not expressed in B cells.) Recover the “T-cell-specific cDNAs” and use them to identify hybridizing DNA clones from a T-cell specific library. Radioactively label those clones that hybridized to the “T-cell-specific” cDNAs, and use cloned probes individually to probe Southern blots of DNA from liver cells and B-cell lymphomas (which would not be expected to rearrange TCR genes) and DNA from different T-cell hybridomas (which would be expected to show different patterns of rearrangements in different T-cell clones). Probing with the TM90 cDNA probe (see bottom of Figure 7-17) resulted in a pattern of bands that did not vary depending on the origin of the cellular DNA. This indicated that TM90 recognizes DNA found in T cells but not B cells, that does not rearrange in different ways in different cells, and therefore is probably not complementary to a TCR gene. In contrast, the TM86 cDNA probe hybridized to bands in the T-cell DNA of different sizes than in either the liver or the B-cell DNA. Furthermore, TM86 showed different patterns of hybridization in different T-cell hybridomas, indicating that the sequence recognized by the TM86 probe is found in varying contexts within the DNA of different hybridomas. Note that this probe also shows the same pattern of bands in the liver cells and B cells. This is consistent with the notion that the genes encoding the TCR recombine in T cells, but not in any other cell type. These results indicated PA R T I I I 250 | TH-cell clone FIGURE 7-17 Production and identification of a cDNA B cell clone encoding the T-cell receptor gene. The flowchart outlines the procedure used by S. M. Hedrick, M. M. Davis, and colleagues to obtain 32P-labeled cDNA clones corresponding to T-cell-specific mRNAs. DNA subtractive hybridization was used to isolate T-cellspecific cDNA fragments that were cloned and labeled with 32P. The labeled cDNA clones were used to probe Southern blots of DNA isolated from embryonic liver (TCR genes should be in the germ-line configuration), from a B-cell lymphoma (TCR genes should also be in the germ-line configuration), and from a panel of T-cell clones. (TCR genes should be differently arranged in each clone.) Their clone, TM86, encodes a gene that rearranged differently in each T-cell clone analyzed. Comparison of its sequence to that of the protein sequence of the TCR isolated by Kappler and Marrack revealed TM86 to encode the chain of the TCR. TM90 cDNA identified the gene for another T cell membrane molecule that does not undergo rearrangement. [Based on S. M. Hedrick et al., 1984, Isolation of cDNA clones encoding mRNA mRNA 97% in free cytoplasmic polyribosomes Adaptive Immunity: Antigen Receptors and MHC 3% in membrane-bound polyribosomes 32P Reverse transcriptase [32P] cDNA T cell-specific membrane-associated proteins, Nature 308:149.] Hybridize A Search for the -Chain Gene Led to the -Chain Gene Instead Hybrids with cDNAs common to T cells and B cells Discard cDNAs specific to T cells Liver cells B cell lymphoma Separate on hydroxyapatite column T-cell clones a b c d e f 10 different cDNA clones Use as probes in Southern blots of genomic DNA Probed with cDNA TM86 Probed with cDNA TM90 that the TM86 cDNA probe was binding to DNA that is rearranged specifically only in T cells and thus most probably encodes a gene for a TCR protein. In this way the first mouse TCR gene was isolated. Comparison of the DNA sequence of Mak’s human clone with the amino acid sequence from a human hybridoma confirmed that Davis and Hedrick et al., and Mak et al., had isolated the gene for the mouse and human forms of the TCR gene, respectively. Focus now switched to the search for the TCR chain. But here, immunology took one of its strange and wonderful twists. Tonegawa’s lab, this time using the subtractive hybridization approach pioneered by Hedrick and Davis, succeeded in cloning a gene that appeared at first to have all the hallmarks of the TCR chain gene. It was expressed in T cells, but not in B cells; it was rearranged in T cell clones; and on sequencing, it revealed regions corresponding to a signal peptide, two Ig family domains, a transmembrane region, and a short cytoplasmic peptide. Furthermore, the predicted molecular weight of the encoded chain appeared to be very close to that of the -chain protein. Tonegawa’s lab initially suggested that “while direct evidence is yet to be produced, it is very likely the pHDS4/203 codes for the subunit of the T cell receptor.” However, biochemical analysis of the TCR and chain proteins by another laboratory had clearly demonstrated that both the and the chains of the TCR heterodimer were glycosylated. A search for potential sequences corresponding to sites of carbohydrate attachment on Tonegawa’s putative -chain sequence came up short. It therefore seemed that the gene Tonegawa’s lab had isolated was not the chain, but something very similar. Further work demonstrated that they had discovered a gene encoding an hitherto unknown receptor that contained both variable and constant regions and recombined only in T cells. They named this gene (and the receptor chain it encoded) . Davis and colleagues next isolated a genomic sequence expressed in T and not B cells that encoded a protein with four potential N-linked glycosylation sites, and a molecular weight consistent with that of the chain of the TCR. With | The Organization and Expression of Lymphocyte Receptor Genes CHAPTER 7 251 Mouse TCR α-chain and δ-chain DNA 5' L Vα1 L Vα2 L Vαn L Vδ 1 L Vδ n Dδ 1 Dδ 2 Jδ 1 Jδ 2 Cδ L Vδ 5 Jα1 Jα2 Jα3 Jαn Cα 3' Mouse TCR β-chain DNA 5' L Vβ 1 L Vβ 2 L Vβn Dβ 1 Jβ 1.1-Jβ 1.7 Cβ 1 Jβ 2.1-Jβ 2.7 Dβ 2 ψ Cβ 2 L Vβ 14 ψ 3' Mouse TCR γ-chain DNA 5' L V γ5 L V γ2 L V γ4 L Vγ 3 J γ1 Cγ1 L 3' Vγ1.3 J γ3 Cγ3 ψ ψ Cγ2 J γ2 L L Jγ4 Cγ4 Vγ1.2 Vγ1.1 = Enhancer ψ = pseudogene FIGURE 7-18 Germ-line organization of the mouse TCR ␣-, ␤-, ␥-, and ␦- chain gene segments. Each C gene segment is composed of a series of exons and introns, which are not shown. The organization of TCR gene segments in humans is similar. Approximate numbers of the gene segments are shown in Table 7-7. [Adapted the genes for the and chains of the TCR fully characterized, those attempting to understand TCR genetics were left with a series of intriguing questions. Was the  chain expressed by the same T cells that expressed the receptor? Did the  chain that Tonegawa had discovered also exist as a heterodimer? If so, what was its partner, and who would find it first? After 3 years of searching, Chien and colleagues, working in the Davis lab, described a fourth TCR gene that they named . Intriguingly, the V chain gene was found to lie downstream of the previously described V genes and just 5 to the J C coding regions. Rearrangement of this locus was found to occur very early in thymic differentiation, and expression of -chain RNA paralleled that of the expression of the  chain in thymic subpopulations, occurring in the same T cells. Most striking about the location of the genes was the observation that, if a T cell rearranged the chain, the intervening genes would be lost (Figure 7-18). Study of the T cells expressing these , receptors demonstrated that , -bearing T cells represented an entirely new T-cell subset TABLE 7-6 Gene Chromosomal locations of TCR genes in humans and mice Human Mouse chain 14 14 chain 7 6  chain 7 13 chain 14 14 from D. Raulet, 1989, The structure, function, and molecular genetics of the gamma/delta T cell receptor, Annual Review of Immunology 7:175; and M. M. Davis, 1990, T cell receptor gene diversity and selection, Annual Review of Biochemistry 59:475.] with an antigen repertoire distinct from that of , T cells and a different anatomical distribution within the lymphoid system in the host. TCR Genes Undergo a Process of Rearrangement Very Similar to That of Ig Genes Table 7-6 shows the chromosomal locations of the TCR receptor genes in mice and humans and Figure 7-18 shows the arrangement of the TCR genes in the mouse genome. The four TCR loci are organized in the germ line in a manner remarkably similar to the multigene organization of the Ig loci, shown in Figure 7-4, and as in the case of Ig genes, functional TCR genes are produced by rearrangements of V and J segments in the - and -chain families and between V, D, and J segments in the - and -chain genes. The approximate numbers of different V, D, and J segments in each of the TCR gene families of mice and humans are shown in Table 7-7. Interestingly, some of the V and V gene segments have been shown to be used in both and TCR chains, which further adds to the diversity of the TCR repertoire. Rearrangement of TCR variable region genes follows the same general outline as that of Ig genes. One point of contrast however is that, whereas Ig light chains incorporate few N nucleotides because of the developmental down-regulation of expression of the TdT enzyme in B cells by the time of light-chain rearrangement, N nucleotides are seen at similar frequencies in all of the TCR chains. The - and -chain variable genes, like the Ig light-chain gene, are generated from one V and one J segment. The TCR and chain variable region genes are assembled from V, 252 TABLE 7-7 PA R T I I I | Adaptive Immunity: Antigen Receptors and MHC TCR gene segments in humans and mice Human Mouse V 40–42 functional; 6 ORFs; 12–13 pseudogenes 21 functional; 1 ORF; 12 pseudogenes D 2 functional 2 functional J 6 J 1; 7 J 2 5 J 1 functional; 2 J 1 ORFs 6 J 2 functional; 1 J 2 pseudogene V 43; some used in conjunction with or constant regions J 50 functional; 8 ORFs; 3 pseudogenes 71 functional; 2 ORFs; 14 pseudogenes 39 functional; 12 ORFs; 10 pseudogenes V 3* 15 functional; 1 pseudogene D 3 2 J 4 2 V 4–6 functional; 3 ORFs; 5 pseudogenes 7 functional J 5 4 * Additional gene segments (5 for human and 9 for mouse) have been found able to associate with downstream or gene segments. A germ-line gene is considered to be functional if the coding region has an ORF without a stop codon and, if there is no described defect in the splicing sites, RSS and/ or regulatory elements. A germ-line entity is considered to be an ORF if the coding region has an open reading frame but alterations have been described in the splicing sites, RSSs, and/or regulatory elements, and/or changes of conserved amino acids have been suggested by the authors to lead to incorrect folding and/or the entity is an orphan. A germ-line entity is considered to be a pseudogene if the coding region has a stop codon(s) and/or a frameshift mutation. In the case of V gene segments, these mutations may be either in the V gene coding sequence, or in the leader sequence. SOURCE: Gene numbers and definitions summarized from the International Immunogenetics Information System Web site: http://imgt.org. Accessed November 16, 2011. D, and J segments, like the Ig heavy chains. In contrast to the Ig genes, however, the TCR families do not have a functionally diverse set of C regions. There is one constant region gene for each of the and gene families, and although there is more than one gene encoding each of the C and C regions, they are highly homologous and do not appear to differ in function. Animals deficient in RAG1/2 or the other enzymes affecting Ig gene recombination, such as Artemis or TdT, are similarly deficient in the generation of both their BCRs and their TCRs, and so it is thought that the essential features of the rearrangement processes are similar in B and T cells (Clinical Focus Box 7-3). The heptamer-nonamer recombination signal sequences are found at the 3 termini of the V sequences, the 5 and 3 termini of the D sequences, and the 5 termini of the J sequences, just as for the Ig variable region gene segments. Recall that the 12-23 rule specifies that recombination can only occur between gene segments contiguous in one case with an RSS containing a 12 (one-turn) bp spacer and in the other with an RSS containing 23 (two-turn) bp spacer (see Figure 7-5). By this rule, recombination between VH and JH immunoglobulin gene segments is disallowed, and hence all properly rearranged heavy-chain V regions must have all three segments represented. In contrast, the D regions of the TCR and chain genes are bordered on the one side by a 12-bp spacer and on the other by a 23-bp spacer (Figure 7-19). By the 12-23 rule, V and J segments from the and gene families can theoretically recombine directly, without an intervening D segment. Furthermore, the rule would also allow for D-D segment recombination. Do these unusual recombinations occur in vivo? In the case of the -chain gene, the answer appears to be that they do not. Evidence obtained from sequencing cDNA fragments containing precisely known V, D, and J sequences has demonstrated that functioning TCR -chain genes always contain one each of the V, D, and J segments. However, the situation is not quite as simple for the TCR -chain genes. Scientists noted that many TCR V genes were considerably longer than TCR V genes; furthermore, the increased sequence length seemed primarily to occur in those parts of the gene encoding the CDR3 residues that make antigen contact. Subsequent analysis showed that many TCR V genes have incorporated not one but two D-region segments, and this additional segment is primarily responsible for the observed increase in gene length. In some V genes, further length was added to the TCR -chain gene by N nucleotide addition at the D-D joint. In contrast to the dramatic levels of diversity facilitated by the incorporation of N nucleotides in all chains of the TCR, TCRs do not appear to undergo somatic hypermutation following antigen contact at any appreciable rate. Table 7-8 compares the mechanisms of the generation and expression of diversity in BCR versus TCR molecules. The Organization and Expression of Lymphocyte Receptor Genes Gene family α β γ δ V region segments D region segments | CHAPTER 7 253 J region segments 5' 5' 5' 3' 3' 3' 5' 5' 5' 3' 3' 3' 5' 5' 5' 3' 3' 3' 5' 5' 5' 3' 3' 3' FIGURE 7-19 The locations of the 12-bp and 23-bp RSS spacers in TCR genes. Note that the spacers on either side of the D and D gene segments are different, allowing for the occurrence of recombined genes bearing two copies of the D . Duplication of D segments has not been observed in vivo. TCR Expression Is Controlled by Allelic Exclusion Allelic exclusion occurs among TCR genes almost as efficiently as among BCR genes, with one exception: some fraction of T cells bear more than one TCR chain. However, it would be unlikely, although not impossible, that a T cell bearing two chains would be able to recognize more than one antigen, given the complexity of T-cell antigen recognition. In Chapter 8, we explain how T cells recognize an antigen made up of a foreign peptide presented to the immune system in a specialized groove in a Major Histocompatibility Complex (MHC) antigen. As we will learn in Chapter 9, TCRs are tested twice within the thymus to determine their functionality before the T cells are released into the periphery. First, T cells are tested to ensure that the and chains can pair into a functioning TCR that is capable of recognizing self MHC antigens bearing self peptides with low to moderate affinity (positive selection). This is necessary because the T cell must have some minimal capacity to recognize self MHC if it is to be able to bind to a peptide presented on the surface of the MHC antigen. Then, T cells bearing receptors that recognize self peptides in association with MHC antigens at high affinity are eliminated in order to eliminate the generation of autoreactive receptors (negative selection). The likelihood that a single TCR chain could pair with more than one TCR chain and meet all these stringent criteria is extremely small. The presence of even one TCR capable of binding self antigens at high affinity would result in the elimination of the cell, irrespective of the specificity of the receptor formed using an alternative chain. Although B and T cells use very similar mechanisms for variable region gene rearrangements, complete rearrangement of Ig genes does not occur in T cells and complete rearrangement of TCR genes does not occur in B cells. Not only is the recombinase enzyme system differently regulated in each lymphocyte lineage, but chromatin is uniquely reconfigured in B cells and T cells to allow the recombinase access only to the appropriate specific antigen receptor genes. TCR Gene Expression Is Tightly Regulated As at Ig loci, expression of TCR genes requires the coordinated use of transcriptional promoters and enhancers that serve at specific stages in T-cell development to make the receptor chromatin open and accessible for recombination (see Chapter 9). T cells begin to rearrange the TCR gene at the pro-T-cell stage. The mouse TCR locus (see Figure 7-18) consists of approximately 40 V genes located upstream of 2 distinct D J clusters. Immediately upstream of each of the two D genes is a promoter region (PD 1 and PD 2). These promoters require the activity of a single enhancer, E , located at the 3end of the locus. Activation of the enhancer serves to facilitate the opening of the chromatin throughout the D J locus, and activation of each of the two PD promoters localizes the rearrangement to their own set of DJ regions. Several TCR enhancer-binding proteins have been identified and, as for B-cell Ig genes, some of these are shared among cells of many different types. Among the T-cellspecific proteins is GATA-3, which binds in a sequencespecific manner to the enhancer elements of all four TCR genes. Tissue- and stage-specific expression of enhancerbinding proteins at the TCR loci helps to facilitate locus accessibility that in turn allows the complex processes of recombination and transcription of the TCR genes and, eventually, cell-surface expression of the TCR proteins. 254 PA R T I I I TABLE 7-8 | Adaptive Immunity: Antigen Receptors and MHC Comparison of the mechanisms of the generation and expression of diversity among B-cell and T-cell receptor molecules Mechanism Used in B cells Used in T cells Comments Multiple germ-line V(D)J genes Yes Yes The mouse V locus has undergone a severe contraction, and, therefore, only 5% of mouse light chains are of the  type. J region diversity is notably higher in TCR -chain genes than in other TCR or Ig genes “Light-chain” segment use  and  variable regions encoded by V and J segments and  variable regions encoded by V and J segments “Heavy-chain” segment use VH regions encoded by V, D and J segments and variable regions encoded by V, D and J segments Absolute dependence on RAG1/2 expression Yes Yes Junctional diversity: P nucleotides and N-nucleotide addition Yes Yes Many fewer N nucleotides found in Ig light chains because of developmental regulation of TdT Multiple D regions per recombined chain Not in Ig heavy chains Present in TCR , but not TCR chains The presence of two D segments allows an additional site for N nucleotide addition. Allelic exclusion of receptor gene expression Absolute Allelic exclusion of TCR genes is not absolute. On activation, secretes product with the same binding site as the receptor Yes TCR is not found in secreted form. Nature of constant region determines function Yes; constant region of secreted antibody product determines its function. No secreted product. Constant region of membrane receptor anchors receptor in membrane and connects with signal transduction complex. Constant region of membrane receptor anchors receptor in membrane and connects with signal transduction complex Receptor genes undergo somatic hypermutation following antigen stimulation Yes No The Organization and Expression of Lymphocyte Receptor Genes | CHAPTER 7 255 BOX 7-3 CLINICAL FOCUS Some Immunodeficiencies Result from Impaired Receptor Gene Recombination Since the RAG1/2 enzymes are responsible for both BCR and TCR gene rearrangements, mutations in the genes encoding RAG1 or RAG2 have catastrophic consequences for the immune system, resulting in patients with severe combined immunodeficiency, or SCID. The RAG1/2 genes are located on human chromosome 11 (i.e., an autosome), so such mutations are inherited in an autosomally recessive manner. Babies born with nonfunctional RAG1 or RAG2 genes have essentially no circulating B or T cells, although they do express normal levels of natural killer (NK) cells and their myeloid and erythroid cells are normal in number and function. Because infants receive antibodies passively from the maternal circulation, the first manifestation of this disease is the complete absence of T-cell function, and hence such infants suffer from severe, recurrent infections with fungi and viruses that would normally be combated by T cells in a healthy neonate. SCID used to be inevitably fatal within the first few months of life, unless babies were delivered directly into a sterile environment. The life span of a SCID patient can be prolonged by preventing contact with all potentially harmful microorganisms. Air must be fil- tered, all food must be sterilized, and there must be no direct contact with other people. Such isolation is feasible only as a temporary measure, pending treatment. Nowadays, babies diagnosed sufficiently early with RAG1/2 deficiency can be successfully treated with bone marrow transplantation. If a suitable donor can be found, and transplantation is performed early in the patient’s first year of life, the chances of the patient surviving to live a normal life are 97% or greater. Patients who carry mutations resulting in partially active or impaired RAG1 or RAG2 genes are diagnosed with Omenn syndrome. Omenn syndrome patients have no circulating B cells and abnormal lymph node architecture with deficiencies in the B-cell zones of the lymph nodes. Although some T cells are present, they are oligoclonal (derived from a very few precursors and hence have very few different receptors), and these T cells tend to be inappropriately activated. They do not respond normally to mitogens or to antigens in vitro. This condition, like RAG1/2 deficiency, is invariably fatal unless corrected by bone marrow transplantation. If deficiency in RAG1/2 genes causes such a catastrophic loss of immune func- tion, it makes sense that loss of any of the other genes implicated in V(D)J recombination would also result in a SCID phenotype. In 1998, it was determined that deficiency of the Artemis DNA repair enzyme also results in the loss of T- and B-cell function, in the face of normal NK cell activity. Loss of Artemis results in a SCID syndrome referred to as Athabascan SCID. Since Artemis is necessary for normal DNA repair, as well as V(D)J recombination, patients with Athabascan SCID also suffer from increased radiation sensitivity of skin fibroblasts and bone marrow cells. Similarly, human patients suffering from a deficiency in DNA ligase IV present with an immunodeficiency that affects T, B, and NK cells. Since DNA ligase IV is implicated in DNA repair functions outside the immune system, such patients also suffer from generalized chromosomal instability, radiosensitivity, and developmental and growth retardation. A SCID defect in mice, first noted and characterized by Prof. Melvin Bosma, results from a nonsense mutation in the gene encoding the DNA-PKcs. Although a similar recessive mutation has been found in Arabian horses, few, if any, human cases of SCID have been reported carrying this mutation. S U M M A R Y ■ ■ ■ The antigen receptor on the surface of a B cell is an Ig with the same binding site as the antibody the B cell’s progeny will eventually secrete. The number of different antibody molecules that can be made by a single individual mouse or human is vast: close to 1013–14. This number is achieved by combinatorial association of multiple gene segments termed V (variable) and J (joining), which together encode the variable region of the light chain of the Ig molecule and V, D (diversity), and J, which encode the variable region of the heavy chain of the Ig molecule. V(D)J recombination occurs at the level of the DNA and is catalyzed by the lymphoid-specific enzymes RAG1 and RAG2. ■ ■ RAG1 and RAG2 recognize Recombination Signal Sequences (RSSs) that are contiguous with the coding regions. RSSs consist of a conserved heptamer and a conserved nonamer sequence, separated by a spacer region conserved in length, but not completely in sequence. Spacer regions have a length corresponding to one turn of the double helix (12-bp spacer) or two turns of the double helix (23-bp spacer). RAG1/2 catalyzes recombination between segments bordered by different spacers. At the sites of recombination, hairpin cleavage can be asymmetric, resulting in P nucleotides, and in the heavychain genes, non-templated nucleotide addition can result in N nucleotides. Exonuclease nibbling can reduce the 256 ■ PA R T I I I | Adaptive Immunity: Antigen Receptors and MHC number of N and P nucleotides in the final receptor gene products. The variable regions of TCRs are generated by the same general mechanism. Most T cells have receptors of the type. The variable regions of chains are made up of V and J segments, whereas those of chains have V, D, and J regions. Other T cells have  receptors. The variable regions of  chains, like those of chains, have V and J segments. chains are encoded within the TCR locus, and their variable regions are made up of V, D, and J segments. but not chains can express more than one D region per chain. R E F E R E N C E S Alder, M. N., et al. 2008. Antibody responses of variable lymphocyte receptors in the lamprey. Nature Immunology 9:319–327. Calame, K., et al. 1980. Mouse Cmu heavy chain immunoglobulin gene segment contains three intervening sequences separating domains. Nature 284:452–455. Capone, M., et al. 1993. TCR beta and TCR alpha gene enhancers confer tissue- and stage-specificity on V(D)J recombination events. EMBO Journal 12:4335–4346. Cesari, I. M., and M. Weigert. 1973. Mouse lambda-chain sequences. Proceedings of the National Academy of Sciences of the United States of America 70:2112–2116. Chien, Y., et al. 1984. A third type of murine T-cell receptor gene. Nature 312:31–35. Chien, Y. H., N. R. Gascoigne, J. Kavaler, N. E. Lee, and M. M. Davis. 1984. Somatic recombination in a murine T-cell receptor gene. Nature 309:322–326. Chun, J. J., D. G. Schatz, M. A. Oettinger, R. Jaenisch, R., and D. Baltimore. 1991. The recombination activating gene-1 (RAG1) transcript is present in the murine central nervous system. Cell 64:189–200. Clatworthy, A. E., M. a. Valencia-Burton, J. E. Haber, and M. A. Oettinger. 2005. The MRE11-RAD50-XRS2 complex, in addition to other non-homologous end-joining factors, is required for V(D)J joining in yeast. Journal of Biological Chemistry 280:20,247–20,252. Cooper, M. D., and M. N. Alder. 2006. The evolution of adaptive immune systems. Cell 124:815–822. Davis, M. M., et al. 1980. An immunoglobulin heavy-chain gene is formed by at least two recombinational events. Nature 283:733–739. Davis, M. M., Y. H. Chien, N. R. Gascoigne, and S. M. Hedrick. 1984. A murine T cell receptor gene complex: Isolation, structure and rearrangement. Immunological Reviews 81:235–258. Early, P., H. Huang, M. Davis, K. Calame, and L. Hood. 1980. An immunoglobulin heavy chain variable region gene is generated from three segments of DNA: VH, D and JH. Cell 19:981–992. Early, P., et al. 1980. Two mRNAs can be produced from a single immunoglobulin mu gene by alternative RNA processing pathways. Cell 20:313–319. Early, P.W., M. M. Davis, D. B. Kaback, N. Davidson, and L. Hood. 1979. Immunoglobulin heavy chain gene organization in mice: Analysis of a myeloma genomic clone containing variable and alpha constant regions. Proceedings of the National Academy of Sciences of the United States of America 76:857–861. Ernst, P., and S. T. Smale. 1995. Combinatorial regulation of transcription II: The immunoglobulin mu heavy chain gene. Immunity 2:427–438. Flajnik, M. F., and M. Kasahara. 2004. Origin and evolution of the adaptive immune system: Genetic events and selective pressures. Nature Reviews Genetics 11:47–59. Gascoigne, N. R., Y. Chien, D. M. Becker, J. Kavaler, and M. M. Davis. 1984. Genomic organization and sequence of T-cell receptor beta-chain constant- and joining-region genes. Nature 310:387–391. Gearhart, P. J. 2004. The birth of molecular immunology. Journal of Immunology 173:4259. Gilfillan, S., A. Dierich, M. Lemeur, C. Benoist, and D. Mathis. 1993. Mice lacking TdT: Mature animals with an immature lymphocyte repertoire. Science 261:1175–1178. Hedrick, S. M., D. I. Cohen, E. A. Nielsen, and M. M. Davis. 1984. Isolation of cDNA clones encoding T cell-specific membrane-associated proteins. Nature 308:149–153. Hedrick, S. M., E. A. Nielsen, J. Kavaler, D. I. Cohen, and M. M. Davis. 1984. Sequence relationships between putative T-cell receptor polypeptides and immunoglobulins. Nature 308:153–158. Herrin, B. R., and M. D. Cooper. 2010. Alternative adaptive immunity in jawless vertebrates. Journal of Immunology 185:1367. Hood, L., et al. 1981. Two types of DNA rearrangements in immunoglobulin genes. Cold Spring Harbor Symposia on Quantitative Biology 45(Pt 2):887–898. Hozumi, N., and S. Tonegawa. 1976. Evidence for somatic rearrangement of immunoglobulin genes coding for variable and constant regions. Proceedings of the National Academy of Sciences of the United States of America 73:3628–3632. Ju, Z., et al. 2007. Evidence for physical interaction between the immunoglobulin heavy chain variable region and the 3 regulatory region. Journal of Biological Chemistry 282:35,169– 35,178. Kasahara, M., T. Suzuki, and L. D. Pasquier. 2004. On the origins of the adaptive immune system: Novel insights from invertebrates and cold-blooded vertebrates. Trends in Immunology 25:105–111. The Organization and Expression of Lymphocyte Receptor Genes Kavaler, J., M. M. Davis, and Y. Chien. 1984. Localization of a T-cell receptor diversity-region element. Nature 310:421–423. Koralov, S. B., T. I. Novobrantseva, J. Konigsmann, A. Ehlich, and K. Rajewsky. 2006. Antibody repertoires generated by VH replacement and direct VH to JH joining. Immunity 25:43–53. Landau, N. R., D. G. Schatz, M. Rosa, and D. Baltimore. 1987. Increased frequency of N-region insertion in a murine pre-B-cell line infected with a terminal deoxynucleotidyl transferase retroviral expression vector. Molecular and Cellular Biology 7:3237–3243. Mansilla-Soto, J., and P. Cortes. 2003. VDJ recombination: Artemis and its in vivo role in hairpin opening. Journal of Experimental Medicine 197:543–547. Oestreich, K. J., et al. 2006. Regulation of TCR beta gene assembly by a promoter/enhancer holocomplex. Immunity 24:381–391. Oettinger, M. A., D. G. Schatz, C. Gorka, and D. Baltimore. 1990. RAG-1 and RAG-2, adjacent genes that synergistically activate V(D)J recombination. Science 248:1517–1523. Pancer, Z., and M. D. Cooper. 2006. The evolution of adaptive immunity. Annual Review of Immunology 24:497–518. Rivera-Munoz, P., et al. 2007. DNA repair and the immune system: From V(D)J recombination to aging lymphocytes. European Journal of Immunology 37(Suppl 1):S71–82. Rooney, S., et al. 2003. Defective DNA repair and increased genomic instability in Artemis-deficient murine cells. Journal of Experimental Medicine 197:553–565. Saito, H., et al. 1984. Complete primary structure of a heterodimeric T-cell receptor deduced from cDNA sequences. Nature 309:757–762. Saito, H., et al. 1984. A third rearranged and expressed gene in a clone of cytotoxic T lymphocytes. Nature 312:36–40. | CHAPTER 7 257 Wei, X. C., et al. 2005. Characterization of the proximal enhancer element and transcriptional regulatory factors for murine recombination activating gene-2. European Journal of Immunology 35:612–621. Weigert, M. G., I. M. Cesari, S. J. Yonkovich, and M. Cohn. 1970. Variability in the lambda light chain sequences of mouse antibody. Nature 228:1045–1047. Zha, S., F. W. Alt, H. L. Cheng, J. W. Brush, and G. Li. 2007. Defective DNA repair and increased genomic instability in Cernunnos-XLF-deficient murine ES cells. Proceedings of the National Academy of Sciences of the United States of America 104:4518–4523. Zhang, Z. 2007. VH replacement in mice and humans. Trends in Immunology 28:132–137. Useful Web Sites http://imgt.org The creators of this website monitor the literature for information on the numbers of genes encoding Ig and TCRs in a variety of species, and regularly update information regarding the numbers of gene segments that have been sequenced. Also includes information on other gene families related to the immune system, such as the MHC. http://cellular-immunity.blogspot.com/2007/12/vdjrecombination.html A clear and brief explanation of V(D)J recombination, with clickable links. http://highered.mcgraw-hill.com/sites/ 0072556781/student_view0/chapter32/animation_ quiz_2.html A clear, even if rather simplistic, animation Schatz, D. G., and Y. Ji. 2011 Recombination centres and the orchestration of V(D)J recombination. Nature Reviews Immunology 11:251–263. of V(D)J recombination, which offers a “big picture” of the process. May be useful to view before launching into more detailed reading. Note that this animation speaks of there being five constant regions, rather than the eight referred to in this chapter; this is because two of the heavy-chain classes— and —can be divided into subclasses, which are ignored in this animation. Schatz, D. G., M. A. Oettinger, and D. Baltimore. 1989. The V(D) J recombination activating gene, RAG-1. Cell 59:1035–1048. http://users.rcn.com/jk imball.ma.ultranet/ BiologyPages/A/AgReceptorDiversity.html A useful Schatz, D. G., and D. Baltimore. 1988. Stable expression of immunoglobulin gene V(D)J recombinase activity by gene transfer into 3T3 fibroblasts. Cell 53:107–115. Tonegawa, S., C. Brack, N. Hozumi, G. Matthyssens, and R. Schuller. 1977. Dynamics of immunoglobulin genes. Immunological Reviews 36:73–94. Tonegawa, S., C. Brack, N. Hozumi, and V. Pirrotta. 1978. Organization of immunoglobulin genes. Cold Spring Harbor Symposia in Quantitative Biology 42(Pt 2):921–931. Tonegawa, S., C. Brack, N. Hozumi, and R. Schuller. 1977. Cloning of an immunoglobulin variable region gene from mouse embryo. Proceedings of the National Academy of Sciences of the United States of America 74:3518–3522. Tonegawa, S., N. Hozumi, G. Matthyssens, and R. Schuller. 1977. Somatic changes in the content and context of immunoglobulin genes. Cold Spring Harbor Symposia in Quantitative Biology 41(Pt 2):877–889. review of antigen receptor diversity with clickable links. Do not be put off by minor differences between this website and this chapter in the numbers of gene segments identified as belonging to particular families—these numbers are constantly being updated, and they vary among different individuals and inbred strains of animals. http://biophilessurf.info/immuno.html A useful collection of databases pertaining to immunological topics, including RAG1 mutations that give rise to immunodeficiencies. www.ncbi.nlm.nih.gov The National Center for Biotechnology Information (NCBI) site offers library tools as well as numerous sequence analysis and protein structure resources (under the “Resources” tab) pertinent to immunology. PA R T I I I 258 S T U D Y | Adaptive Immunity: Antigen Receptors and MHC Q U E S T I O N S 1. Indicate whether each of the following statements is true or false. If you think a statement is false, explain why. a. In generating a B-cell receptor gene, V segments some- times join to C segments. b. In generating a T-cell receptor gene, V segments some- times join to C . c. The switch in constant region use from to IgM to IgD is mediated by DNA rearrangements. d. Although each B cell carries two alleles encoding the immunoglobulin heavy and light chains, only one allele is expressed. e. Like the variable regions of the heavy chain of the B-cell Ig receptor, TCR variable genes are all encoded in three segments. 2. Explain why a VH segment cannot join directly with a JH segment in heavy-chain gene rearrangement. 3. For each incomplete statement below, select the phrase(s) that correctly completes the statement. More than one choice may be correct. a. Recombination of Ig gene segments serves to: (1) Promote Ig diversification (2) Assemble a complete Ig coding sequence (3) Allow changes in coding information during B-cell maturation (4) Increase the affinity of Ig for antibody (5) All of the above b. Kappa and lambda light-chain genes: (1) Are located on the same chromosome (2) Associate with only one type of heavy chain (3) Can be expressed by the same B cell (4) All of the above (5) None of the above c. Generation of combinatorial diversity within the vari- able regions of Ig involves mRNA splicing DNA rearrangement Recombination signal sequences One-turn/two-turn joining rule Switch sites d. The mechanism that permits Ig to be synthesized in either a membrane-bound or secreted form is: (1) Allelic exclusion (2) Codominant expression (3) Class switch recombination (4) The one-turn/two-turn joining rule (5) Differential RNA processing e. During Ig VH recombination, the processes that contribute to additional diversity at the third complementarity-determining region of Ig variable regions include: (1) Introduction of the DH gene segments into the heavy-chain V gene (2) mRNA splicing of the membrane form of the C region of the constant chain (3) Exonuclease cleavage of the ends of the gene segments (4) P nucleotide addition (5) N nucleotide addition (1) (2) (3) (4) (5) 4. You have been given a cloned mouse myeloma cell line that secretes IgG with the molecular formula 22. Both the heavy and light chains in this cell line are encoded by genes derived from allele 1 (i.e., the first of the two homologous alleles encoding each type of chain). Indicate the form(s) in which each of the genes listed below would occur in this cell line using the following symbols: G  germ-line form; P  productively rearranged form; NP  nonproductively rearranged form. State the reason for your choice in each case. (a) Heavy-chain allele 1 (b) Heavy-chain allele 2 (c) -chain allele 1 (d) -chain allele 2 (e) -chain allele 1 (f) -chain allele 2 5. You have identified a B-cell lymphoma that has made non- productive rearrangements for both heavy-chain alleles. What is the arrangement of its light-chain DNA? Why? 6. The random addition of nucleotides by TdT is a wasteful and potentially risky evolutionary strategy. State why it may be disadvantageous to the organism and why, therefore, you think it is sufficiently useful to have been retained during vertebrate evolution. 7. Are there any differences in the genetic strategies used to generate complete V genes in T and BCRs? 8. What known features of the TCR did Hedrick, Davis, and colleagues use in their quest to isolate the TCR genes? 9. The following figure describes the end of a V region sequence and the beginning of the D region sequence to which is it about to be joined. Arrows mark the cleavage points where the RAG1/RAG2 complex will make the cut and recombination will be targeted. V 5' 3' AGCATC TATCGA TCGTAG ATAGCT D 3' 5' a. Is this a heavy-chain or a light-chain sequence? How do you know? b. What DNA sequence structure would you find just downstream of the AGCATC sequence immediately adjacent to the 3 end of the V segment? c. Here is one possible VD joint structure formed after recombination between these two gene segments: 5' 3' V AGCATCGACGCCGTATCGA D TCGTAGCTGCGGCATAGCT (1) Which residue(s) MAY be P-region nucleotide(s)? (2) Can we know for certain that this residue is a P-region nucleotide? (3) Which residues must have been added by TdT, and therefore must be N-region nucleotides? (4) Can we know for certain whether a residue is an N-region nucleotide? The Organization and Expression of Lymphocyte Receptor Genes 10. Challenge data question. In Classic Experiment Box 7.1, Figure 1, we see the pattern of bands that Hozumi and Tonegawa’s experiment would have revealed, using the particular cell line that they used. Below, see a pair of gels that represent the results of a hypothetical experiment performed using the same general protocol. In this hypothetical experiment, our probes correspond to either the V region or the C region. Furthermore, the investigators used a different tumor cell line and a different restriction endonuclease. Germline DNA DNA from antibody producing cells C-region probe V-region probe C-region probe V-region probe 1 2 3 4 | CHAPTER 7 259 a. Why are there two C-region bands in the germ-line blot? b. Why are these two bands still present in the myeloma blot, and why are there two bands recognized by the V-region probe on the myeloma blot? c. From this plot, would you hypothesize that the cell had achieved a successful arrangement at the first  allele? Why or why not? d. How would you prove your answer to part c? This page lelt intentionally blank. 8 The Major Histocompatibility Complex and Antigen Presentation A lthough both T and B cells use surface molecules to recognize antigen, they accomplish this in very different ways. In contrast to antibodies or B-cell receptors, which can recognize an antigen alone, T-cell receptors only recognize pieces of antigen that are positioned on the surface of other cells. These antigen pieces are held within the binding groove of a cell surface protein called the major histocompatibility complex (MHC) molecule, encoded by a cluster of genes collectively called the MHC locus. These fragments are generated inside the cell following antigen digestion, and the complex of the antigenic peptide plus MHC molecule then appears on the cell surface. MHC molecules thus act as a cell surface vessel for holding and displaying fragments of antigen so that approaching T cells can engage with this molecular complex via their T-cell receptors. The MHC got its name from the fact that the genes in this region encode proteins that determine whether a tissue transplanted between two individuals will be accepted or rejected. The pioneering work of Benacerraf, Dausset, and Snell helped to characterize the functions controlled by the MHC, specficially, organ transplant fate and the immune responses to antigen, resulting in the 1980 Nobel Prize in Medicine and Physiology for the trio (see Table 1-2). Follow-up studies by Rolf Zinkernagel and Peter Doherty illustrated that the proteins encoded by these genes play a seminal role in adaptive immunity by showing that T cells recognize MHC proteins as well as antigen. Structural studies done by Don Wiley and others showed that different MHC proteins bind and present different antigen fragments. There are many alleles of most MHC genes, and the specific alleles one inherits play a significant role in susceptibility to disease, including the development of autoimmunity. The mechanisms by which this family of molecules exerts such a strong influence on the development of immunity to nearly all types of antigens has become a major theme in immunology, and has taken Ribbon diagram of a human MHC class I molecule (blue) with a space-filled peptide (orange) held in the binding groove. ■ The Structure and Function of MHC Molecules ■ General Organization and Inheritance of the MHC ■ The Role of the MHC and Expression Patterns ■ ■ The Endogenous Pathway of Antigen Processing and Presentation The Exogenous Pathway of Antigen Processing and Presentation ■ Cross-Presentation of Exogenous Antigens ■ Presentation of Nonpeptide Antigens the study of the MHC far beyond its origins in the field of transplantation biology. There are two main classes of MHC molecules: class I and class II. These two molecules are very similar in their final quaternary structure, although they differ in how they create these shapes via primary through quaternary protein arrangements. Class I and class II MHC molecules also differ in terms of which cells express them and in the source of the antigens they present to T cells. Class I molecules are present on all nucleated cells in the body and specialize in presenting antigens that originate from 261 262 PA R T I I I | Adaptive Immunity: Antigen Receptors and MHC the cytosol, such as viral proteins. These are presented to CD8⫹ T cells, which recognize and kill cells expressing such intracellular antigens. In contrast, class II MHC molecules are expressed almost exclusively on a subset of leukocytes called antigen-presenting cells (APCs) and specialize in presenting antigens from extracellular spaces that have been engulfed by these cells, such as fungi and extracellular bacteria. Once expressed on the cell surface, the MHC class II molecule presents the antigenic peptide to CD4⫹ T cells, which then become activated and go on to stimulate immunity directed primarily toward destroying extracellular invaders. This chapter will begin by discussing the structure of MHC molecules, followed by the genetic organization of the DNA region that encodes these proteins and inheritance patterns. We then turn to the function of the MHC in regulating immunity, including reference to several seminal studies in these areas. In this section we also discuss regulation of MHC expression. This is followed by a detailed discussion of the cellular pathways that lead to antigen degradation and association with each type of MHC molecule (antigen processing) and the appearance of these MHC-peptide complexes on the cell surface for recognition by T cells (antigen presentation). The role of particular APCs in these processes will also be presented (no pun intended!). The chapter concludes with a discussion of unique processing and presentation pathways, such as cross-presentation and the handling of nonpeptide antigens by the immune system. The Structure and Function of MHC Molecules Class I and class II MHC molecules are membrane-bound glycoproteins that are closely related in both structure and function. Both classes of MHC molecule have been isolated and purified, and the three-dimensional structures of their extracellular domains have been resolved by x-ray crystallography. These membrane glycoproteins function as highly specialized antigen-presenting molecules with grooves that form unusually stable complexes with peptide ligands, displaying them on the cell surface for recognition by T cells via T-cell receptor (TCR) engagement. In contrast, class III MHC molecules are a group of unrelated proteins that do not share structural similarity or function with class I and II molecules, although many of them do participate in other aspects of the immune response. Class I Molecules Have a Glycoprotein Heavy Chain and a Small Protein Light Chain Two polypeptides assemble to form a single class I MHC molecule: a 45-kilodalton (kDa)  chain and a 12-kDa 2-microglobulin molecule (Figure 8-1, left). The  chain is organized into three external domains (1, 2, and 3), each approximately 90 amino acids long; a transmembrane domain of about 25 hydrophobic amino acids followed by a short stretch of charged (hydrophilic) amino acids; and a cytoplasmic anchor segment of 30 amino acids. Its companion, 2-microglobulin, is similar in size and organization to the 3 domain. 2-microglobulin does not contain a transmembrane region and is noncovalently bound to the MHC class I  chain. Sequence data reveal strong homology between the 3 domain of MHC class I, 2-microglobulin, and the constant-region domains found in immunoglobulins. The 1 and 2 domains interact to form a platform of eight antiparallel  strands spanned by two long -helical regions (Figure 8-2a). The structure forms a deep groove, or cleft, with the long  helices as sides and the  strands of the  sheet as the bottom (Figure 8-2b). This peptide-binding groove is located on the top surface of the class I MHC molecule, and it is large enough to bind a peptide of 8 to 10 amino acids. During the x-ray crystallographic analysis of class I molecules, small noncovalently associated peptides that had co-crystallized with the protein were found in the groove. The big surprise came when these peptides were later discovered to be processed self-proteins bound to this deep groove and not the foreign antigens that were expected. The 3 domain and 2-microglobulin are organized into two  pleated sheets each formed by antiparallel  strands of amino acids. As described in Chapter 3, this structure, known as the immunoglobulin fold, is characteristic of immunoglobulin domains. Because of this structural similarity, which is not surprising given the considerable sequence similarity with the immunoglobulin constant regions, class I MHC molecules and 2-microglobulin are classified as members of the immunoglobulin superfamily (see Figure 3-19). The 3 domain appears to be highly conserved among class I MHC molecules and contains a sequence that interacts strongly with the CD8 cell surface molecule found on TC cells. All three molecules (class I  chain, 2-microglobulin, and a peptide) are essential to the proper folding and expression of the MHC-peptide complex on the cell surface. This is demonstrated in vitro using Daudi tumor cells, which are unable to synthesize 2-microglobulin. These tumor cells produce class I MHC  chains but do not express them on the cell membrane. However, if Daudi cells are transfected with a functional gene encoding 2-microglobulin, class I molecules will appear on the cell surface. Class II Molecules Have Two Non-Identical Glycoprotein Chains Class II MHC molecules contain two different polypeptide chains, a 33-kDa  chain and a 28-kDa  chain, which associate by noncovalent interactions (see Figure 8-1, right). Like class I  chains, class II MHC molecules are membranebound glycoproteins that contain external domains, a transmembrane segment, and a cytoplasmic anchor segment. Each The Major Histocompatibility Complex and Antigen Presentation | CHAPTER 8 Class I molecule α2 Class II molecule α1 Peptide-binding groove α1 Membrane - distal domains S S S α3 S β1 S S S Membrane -proximal domains (Ig-fold structure) 263 S S β2-microglobulin α2 S S S β2 Transmembrane segment Cytoplasmic tail FIGURE 8-1 Schematic diagrams of class I and class II MHC molecules showing the external domains, transmembrane segments, and cytoplasmic tails. The peptide-binding groove is formed by the membrane-distal domains in both class I and class II molecules (blue). The membrane-proximal domains (green and red) possess the basic immunoglobulin-fold structure; thus, both class I and class II MHC molecules are classified as members of the immunoglobulin superfamily. chain in a class II molecule contains two external domains: 1 and 2 domains in one chain and 1 and 2 domains in the other (Figure 8-2c). The membrane-proximal 2 and 2 domains, like the membrane-proximal 3/2-microglobulin domains of class I MHC molecules, bear sequence similarity to the immunoglobulin-fold structure. For this reason, class II MHC molecules are also classified in the immunoglobulin superfamily. The 1 and 1 domains form the peptide-binding groove for processed antigen (Figure 8-2d). Although similar to the peptide-binding groove of MHC class I, this groove is thus formed by the association of two separate chains. X-ray crystallographic analysis reveals the similarity between these two classes of molecule, strikingly apparent when class I and class II molecules are superimposed (Figure 8-3). Interestingly, despite the fact that these two structures are encoded quite differentially (one chain versus two), the final quaternary structure is similar and retains the same overall function: the ability to bind antigen and present it to T cells. The peptide-binding groove of class II molecules, like that found in class I molecules, is composed of a floor of eight antiparallel  strands and sides of antiparallel  helices, where peptides typically ranging from 13 to 18 amino acids can bind. The class II molecule lacks the conserved residues in the class I molecule that bind to the terminal amino acids of short peptides, and therefore forms more of an open pocket. In this way, class I presents more of a socket-like opening, whereas class II pos- sesses an open-ended groove. The functional consequences of these differences in fine structure will be explored below. Class I and II Molecules Exhibit Polymorphism in the Region That Binds to Peptides Several hundred different allelic variants of class I and II MHC molecules have been identified in humans. Any one individual, however, expresses only a small number of these molecules—up to six different class I molecules and 12 or more different class II molecules. Yet this limited number of MHC molecules must be able to present an enormous array of different antigenic peptides to T cells, permitting the immune system to respond specifically to a wide variety of antigenic challenges. Thus, peptide binding by class I and II molecules does not exhibit the fine specificity characteristic of antigen binding by antibodies and T-cell receptors. Instead, a given MHC molecule can bind numerous different peptides, and some peptides can bind to several different MHC molecules. Because of this broad specificity, the binding between a peptide and an MHC molecule is often referred to as “promiscuous.” This promiscuity of peptide binding allows many different peptides to match up with the MHC binding groove and for exchange of peptides to happen on occasion, unlike the relatively stable, high-affinity cognate interactions of antibodies and T-cell receptors with their specific ligands. 264 PA R T I I I | Adaptive Immunity: Antigen Receptors and MHC (a) MHC class I (b) MHC class I peptide-binding groove Peptide-binding groove α1 domain α1 domain α helix α2 domain β sheets α2 domain β2-microglobulin α3 domain (c) MHC class II (d) MHC class II peptide-binding groove Peptide-binding groove α1 domain α helix β1 domain α1 domain β sheets β1 domain β2 domain α2 domain FIGURE 8-2 Representations of the three-dimensional structure of the external domains of human MHC class I and class II molecules based on x-ray crystallographic analysis. (a, c) Side views of class I and II, respectively, in which the  strands are depicted as thick arrows and the  helices as spiral ribbons. Disulfide bonds are shown as two interconnected spheres. The 1 and 2 domains of class I and the 1 and 1 domains of class II interact to form the peptide-binding groove. Note the immunoglobulin-fold structure of each membrane proximal domain, including the 2-microglobulin molecule. (b, d) The 1 (dark blue) and 2 domains (light blue) of class I and the 1 (dark blue) and 1 (light blue) domains of class II as viewed from the top, showing the peptide-binding grooves, consisting of a base of antiparallel  strands and sides of  helices for each molecule. This groove can accommodate peptides of 8 to 10 residues for class I and 13 to 18 residues for class II. The Major Histocompatibility Complex and Antigen Presentation | CHAPTER 8 265 no such requirement for class II peptide binding. The association of peptide and MHC molecule is very stable under physiologic conditions (Kd values range from ~ 106 to 1010). Thus, most of the MHC molecules expressed on the membrane of a cell will be associated with a peptide. FIGURE 8-3 The peptide-binding groove of a human class II MHC molecule (blue), superimposed over the corresponding regions of a human class I MHC molecule (red). Overlapping regions are shown in white. These two molecules create very similar binding grooves such that most of the structural differences (long stretches, kinks, bends, etc.) lie outside the peptide-binding regions. [From J. H. Brown et al., 1993, Nature 364:33.] Given the similarities in the structures of the peptidebinding grooves in class I and II MHC molecules, it is not surprising that they exhibit some common peptide-binding features (Table 8-1). In both types of MHC molecules, peptide ligands are held in a largely extended conformation that runs the length of the groove. The peptide-binding groove in class I molecules is blocked at both ends, whereas the ends of the groove are open in class II molecules (Figure 8-4). As a result of this difference, class I molecules bind peptides that typically contain 8 to 10 amino acid residues, whereas the open groove of class II molecules accommodates slightly longer peptides of 13 to 18 amino acids. Another difference is that class I binding requires that the peptide contain specific amino acid residues near the N and C termini; there is TABLE 8-1 Class I MHC-Peptide Interaction Class I MHC molecules bind peptides and present them to CD8⫹ T cells. These peptides are often derived from endogenous intracellular proteins that are digested in the cytosol. The peptides are then transported from the cytosol into the endoplasmic reticulum (ER), where they interact with class I MHC molecules. This process, known as the cytosolic or endogenous processing pathway, is discussed in detail later in this chapter. Each human or mouse cell can express several unique class I MHC molecules, each with slightly different rules for peptide binding. Because a single nucleated cell expresses about 105 copies of each of these unique class I molecules, each with its own peptide promiscuity rules, many different peptides will be expressed simultaneously on the surface of a nucleated cell. This means that although many of the peptide fragments of a given foreign antigen will be presented to CD8⫹ T cells, the set of MHC class I alleles inherited by each individual will determine which specific peptide fragments from a larger protein get presented. The bound peptides isolated from different class I molecules have two distinguishing features: they are 8 to 10 amino acids in length, most commonly 9, and they contain specific amino acid residues at key locations in the sequence. The ability of an individual class I MHC molecule to bind to a diverse spectrum of peptides is due to the presence of the same or similar amino acid residues at these key positions along the peptides (Figure 8-5). Because these amino acid residues anchor the peptide into the groove of the MHC molecule, they are called anchor residues. The side chains of the anchor residues in the peptide are complementary with surface features of the binding groove of the class I MHC molecule. The amino acid residues lining the binding sites vary among different class I allelic variants, and this is what determines the chemical Peptide binding by class I and class II MHC molecules Class I molecules Class II molecules Peptide-binding domain 1/2 1/1 Nature of peptide-binding groove Closed at both ends Open at both ends General size of bound peptides 8–10 amino acids 13–18 amino acids Peptide motifs involved in binding to MHC molecule Anchor residues at both ends of peptide; generally hydrophobic carboxyl-terminal anchor Conserved residues distributed along the length of the peptide Nature of bound peptide Extended structure in which both ends interact with MHC groove but middle arches up away from MHC molecule Extended structure that is held at a constant elevation above the floor of the MHC groove 266 (a) PA R T I I I | Adaptive Immunity: Antigen Receptors and MHC α1 domain Eluted from MHC protein 1 Eluted from MHC protein 2 1 2 3 4 5 6 7 8 9 H3 N V G P Q K N E N L COO− H3 N S G P R K A I A L COO− H3N V G P S G K Y F I COO− H3N S G P E R I L S L COO− H3 N S Y F P E I T H I COO− H3N T Y Q R T R A L V COO− H3N S Y I G S I N N I COO− A = alanine E = glutamic acid F = phenylalanine G = glycine H = histidine I = isoleucine α2 domain (b) α1 domain K = lysine L = leucine N = asparagine P = proline Q = glutamine R = arginine S = serine T = threonine V = valine Y = tyrosine FIGURE 8-5 Examples of anchor residues (blue) in nonameric peptides eluted from two different class I MHC molecules. Anchor residues, at positions 2/3 and 9, that interact with the class I MHC molecule tend to be hydrophobic amino acids. The two MHC proteins bind peptides with different anchor residues. [Data from V. H. Engelhard, 1994, Current Opinion in Immunology 6:13.] β1 domain FIGURE 8-4 MHC class I and class II molecules with bound peptides. (a) Ribbon model of human class I molecule HLA-A2 (1 in dark blue, 2 in light blue) with a peptide from the HIV-1 envelope protein GP120 (space-filling, orange) in the binding groove. (b) Ribbon model of human class II molecules HLA-DR1 with the DR chain shown in dark blue and the DR chain in light blue. The peptide (space-filling, orange) in the binding groove is from influenza hemagglutinin (amino acid residues 306–318). [Part (a) PDB ID 1HHG, part (b) PDB ID 1DLH.] identity of the anchor residues that can interact with a given class I molecule. In a critical study of peptide binding by MHC molecules, peptides bound by two different class I MHC proteins were released chemically and analyzed by high-performance liquid chromatography (HPLC) mass spectrometry. More than 2000 distinct peptides were found among the peptide ligands released from these two class I MHC molecules. Since there are approximately 105 copies of each class I protein per cell, it is estimated that each of the 2000 distinct peptides is presented with a frequency of between 100 and 4000 copies per cell. Evidence suggests that even a single MHC-peptide complex may be sufficient to target a cell for recognition and lysis by a cytotoxic T lymphocyte with a receptor specific for that target structure. All peptides examined to date that bind to class I molecules contain a carboxyl-terminal anchor (e.g., see position 9 in Figure 8-5). These anchors are generally hydrophobic residues (e.g., leucine, isoleucine), although a few charged amino acids have been reported. Besides the anchor residue found at the carboxyl terminus, another anchor is often found at the second, or second and third, positions at the amino-terminal end of the peptide. In general, any peptide of the correct length that contains the same or chemically similar anchor residues will bind to the same class I MHC molecule. Knowledge of these key positions and the chemical restrictions for amino acids at these positions may allow for more nuanced clinical design studies, such as future vaccines targeted at eliciting protective immunity to particular pathogens. X-ray crystallographic analyses of class I MHC-peptide complexes have revealed how the peptide-binding groove in a given MHC molecule can interact stably with a broad spectrum of different peptides. The anchor residues at both ends of the peptide are buried within the binding groove, thereby holding the peptide firmly in place (see Figure 8-4a). Nonameric peptides are bound preferentially. The main contacts between class I MHC molecules and the peptides they bind The Major Histocompatibility Complex and Antigen Presentation Bulge 4 5 6 N 1 2 3 7 8 9 C Hydrogen bonds with MHC molecule FIGURE 8-6 Conformation of peptides bound to class I MHC molecules. Schematic diagram depicting, in a side view, the conformational variance in MHC class I-bound peptides of different lengths. Longer peptides bulge in the middle, presumably interacting more with TCR, whereas shorter peptides lay flat in the groove. Contact with the MHC molecule is by hydrogen bonds to anchor residues 2 and/or 3 and 9. [Adapted from P. Parham, 1992, Nature 360:300, | CHAPTER 8 267 but the binding characteristics are determined by the central 13 residues. The peptides that bind to a particular class II molecule often have internal conserved sequence motifs, but unlike class I–binding peptides, they appear to lack conserved anchor residues (see Table 8-1). Instead, hydrogen bonds between the backbone of the peptide and the class II molecule are distributed throughout the binding site rather than being clustered predominantly at the ends of the site as is seen in class I–bound peptides. Peptides that bind to class II MHC molecules contain an internal sequence of 7 to 10 amino acids that provide the major contact points. Generally, this sequence has an aromatic or hydrophobic residue at the amino terminus and three additional hydrophobic residues in the middle portion and carboxyl-terminal end of the peptide. In addition, over 30% of the peptides eluted from class II molecules contain a proline residue at position 2 and another cluster of prolines at the carboxyl-terminal end. This relative flexibility contributes to peptide binding promiscuity. © 1992 Macmillan Magazines Limited] involve residue 2 at the amino-terminal end and residue 9 at the carboxyl terminus of the peptide. Between the anchors, the peptide arches away from the floor of the groove in the middle (Figure 8-6), allowing peptides that are slightly longer or shorter to be accommodated. Amino acids in the center of the peptide that arch away from the MHC molecule are more exposed and presumably can interact more directly with the T-cell receptor. Class II MHC-Peptide Interaction Class II MHC molecules bind and present peptides to CD4⫹ T cells. Like class I molecules, MHC class II molecules can bind a variety of peptides. These peptides are typically derived from exogenous proteins (either self or nonself) that are degraded within the exogenous processing pathway, described later in this chapter. Most of the peptides associated with class II MHC molecules are derived from self membrane-bound proteins or foreign proteins internalized by phagocytosis or by receptor-mediated endocytosis and then processed through the exogenous pathway. Peptides recovered from class II MHC-peptide complexes generally contain 13 to 18 amino acid residues, somewhat longer than the nonameric peptides that most commonly bind to class I molecules. The peptide-binding groove in class II molecules is open at both ends (see Figure 8-4b), allowing longer peptides to extend beyond the ends, like an extra-long hot dog in a bun. Peptides bound to class II MHC molecules maintain a roughly constant elevation on the floor of the binding groove, another feature that distinguishes peptide binding to class I and class II molecules. Peptide-binding studies and structural data for class II molecules indicate that a central core of 13 amino acids determines the ability of a peptide to bind class II. Longer peptides may be accommodated within the class II groove, General Organization and Inheritance of the MHC Every vertebrate species studied to date possesses the tightly linked cluster of genes that constitute the MHC. As we have just discussed, MHC molecules have the important job of deciding which fragments of a foreign antigen will be “seen” by the host T cells. In general terms, MHC molecules face a similar ligand binding challenge to that faced, collectively, by B-cell and T-cell receptors: they must be able to bind a wide variety of antigens, and they must do so with relatively strong affinity. However, these immunologically relevant molecules meet this challenge using very different strategies. Although B- and T-cell receptor diversity is generated through genomic rearrangement and gene editing (described in Chapter 7), MHC molecules have opted for a combination of peptide binding promiscuity (discussed above) and the expression of several different MHC molecules on every cell (described below). Using this clever combined strategy, the immune system has evolved a way of maximizing the chances that many different regions, or epitopes, of an antigen will be recognized. Studies of the MHC gene cluster originated when it was found that the rejection of foreign tissue transplanted between individuals in a species was the result of an immune response mounted against cell surface molecules, now called histocompatibility antigens. In the mid-1930s, Peter Gorer, who was using inbred strains of mice to identify blood-group antigens, identified four groups of genes that encode blood-cell antigens. He designated these I through IV. Work carried out in the 1940s and 1950s by Gorer and George Snell established that antigens encoded by the genes in the group designated as II took part in the rejection of transplanted tumors and other tissue. Snell called these histocompatibility genes; their current designation as histocompatibility-2 (H-2, or MHC) genes in the mouse was in reference to Gorer’s original group II blood-cell antigens. 268 PA R T I I I | Adaptive Immunity: Antigen Receptors and MHC Mouse H-2 complex Complex H–2 II MHC class I Region K IA IE H–2K IA αβ IE αβ Gene products III I S D TNF-α Lymphotoxin-α C′ proteins H–2D H–2L* * Not present in all haplotypes Human HLA complex HLA Complex II MHC class I III Region DP DQ DR Gene products DP αβ DQ αβ DR αβ C4, C2, BF C′ proteins TNF-α Lymphotoxin-α B C A HLA-B HLA-C HLA-A FIGURE 8-7 Comparison of the organization of the major histocompatibility complex (MHC) in mouse and human. The MHC is referred to as the H-2 complex in mice and as the HLA complex in humans. In both species, the MHC is organized into a number of regions encoding class I (pink), class II (blue), and class III (green) gene products. The class I and II gene products shown in this figure are considered to be the classical MHC molecules. The class III gene products include other immune function–related compounds such as the complement proteins (C’) and tumor necrosis factors (TNF- and Lymphotoxin ). The MHC Locus Encodes Three Major Classes of Molecules human, this region in mouse is noncontinuous, interrupted by the class II and III regions (Figure 8-8). Recall that there are two chains to the MHC class I molecule: the more variable and antigen-binding  chain and the common -2-microglobulin chain. The  chain molecules are encoded by the K and D regions in mice, with an additional L region found in some strains, and by the A, B, and C loci in humans. 2-microglobulin is encoded by a gene outside the MHC. Collectively, these are referred to as classical class I molecules; all posses the functional capability of presenting protein fragments of antigen to T cells. Additional genes or groups of genes within the class I region of both mouse and human encode nonclassical class I molecules that are expressed only in specific cell types and have more specialized functions. Some appear to play a role in self/nonself discrimination. One example is the class I HLA-G molecule. These are present on fetal cells at the maternal-fetal interface and are credited with inhibiting rejection by maternal CD8⫹ T cells by protecting the fetus from identification as foreign, which may occur when paternally derived antigens begin to appear on the developing fetus. Class II MHC molecules are encoded by the IA and IE regions in mice and by the DP, DQ, and DR regions in humans (see Figures 8-7 and 8-8). The terminology is somewhat confusing, since the D region in mice encodes class I MHC molecules, whereas DP, DQ, and DR in humans refers to class II genes and molecules. Recall that class II molecules are composed of two chains, both of which interact with The major histocompatibility complex is a collection of genes arrayed within a long continuous stretch of DNA on chromosome 6 in humans and on chromosome 17 in mice. The MHC is referred to as the human leukocyte antigen (HLA) complex in humans and as the H-2 complex in mice, the two species in which these regions have been most studied. Although the arrangement of genes is somewhat different in the two species, in both cases the MHC genes are organized into regions encoding three classes of molecules (Figure 8-7): • Class I MHC genes encode glycoproteins expressed on the surface of nearly all nucleated cells; the major function of the class I gene products is presentation of endogenous peptide antigens to CD8⫹ T cells. • Class II MHC genes encode glycoproteins expressed predominantly on APCs (macrophages, dendritic cells, and B cells), where they primarily present exogenous antigenic peptides to CD4⫹ T cells. • Class III MHC genes encode several different proteins, some with immune functions, including components of the complement system and molecules involved in inflammation. Class I MHC molecules were the first discovered and are expressed in the widest range of cell types. Unlike in the The Major Histocompatibility Complex and Antigen Presentation | CHAPTER 8 269 MOUSE CHROMOSOME 17 H-2 Complex 1500 kb D L Telomere C′ IE α LMP2 TAP2 LMP7 TAP1 Oβ IAβ IA α IEβ I TNF- α Lymphotoxin-α III II Oα Mα Mβ Loci I K Class Centromere HUMAN CHROMOSOME 6 HLA 4000 kb KEY Gene LMP2, LMP7 TAP1, TAP2 TNF- α, Lymphotoxin- α HLA-G HLA-E HLA-B HLA-C HLA-F Telomere Centromere I TNF- α Lymphotoxin-α C′ DRα III DRβ DP β DP α Loci DO α DM α DMβ LMP2 TAP1 LMP7 TAP2 DOβ DQβ DQα II Class HLA-A Complex Encoded protein Proteasome-like subunits Peptide-transporter subunits Tumor necrosis factor α and lymphotoxin β FIGURE 8-8 Simplified map of the mouse and human MHC loci. The MHC class I genes are colored red, MHC class II genes are colored blue, and genes in MHC III are colored green. Classical class I genes are labeled in red, classical class II genes are labeled in blue, and the nonclassical MHC genes are labeled in black. The concept of classical and nonclassical does not apply to class III. Only some of the non-classical class I genes are shown; the functions of only some of their proteins are known. antigen. The class II region of the MHC locus encodes both the  chain and the  chain of a particular class II MHC molecule, and in some cases multiple genes are present for either or both chains. For example, individuals can inherit up to four functional DR -chain genes, and all of these are expressed simultaneously in the cell. This allows any DR -chain gene product to pair with any DR  chain product. Since the antigen-binding groove of class II is formed by a combination of the  and  chains, this creates several unique antigen-presenting DR molecules on the cell. As with the class I loci, additional nonclassical class II molecules with specialized immune functions are encoded within this region. For instance, human nonclassical class II genes designated DM and DO have been identified. The DM genes encode a class II–like molecule (HLA-DM) that facilitates the loading of antigenic peptides into class II MHC molecules. Class II DO molecules, which are expressed only in the thymus and on mature B cells, have been shown to serve as regulators of class II antigen processing. Class I and II MHC molecules have common structural features, and both have roles in antigen processing and presentation. By contrast, the class III MHC region encodes several molecules that are critical to immune function but have little in common with class I or II molecules. Class III products include the complement components C4, C2, and factor B (see Chapter 6), as well as several inflammatory cytokines, including the two tumor necrosis factor proteins (TNF- and Lymphotoxin- [TNF-]). Allelic variants of some of these class III MHC gene products have been linked to certain diseases. For example, polymorphisms within the TNF- gene, which encodes a cytokine involved in many immune processes (see Chapter 4), have been linked to susceptibility to certain infectious diseases and some forms of autoimmunity, including Crohn’s disease and rheumatic arthritis. Despite its differences from class I and class II genes, the linked class III gene cluster is conserved in all species with an MHC region, suggesting similar evolutionary pressures have come to bear on this cluster of genes. PA R T I I I 270 | Adaptive Immunity: Antigen Receptors and MHC The Exon/Intron Arrangement of Class I and II Genes Reflects Their Domain Structure Separate exons encode each region of the class I and II proteins. Each of the mouse and human class I genes has a 5 leader exon encoding a short signal peptide followed by five or six exons encoding the  chain of the class I molecule (Figure 8-9a). The signal peptide serves to facilitate insertion of the  chain into the ER and is removed by proteolytic enzymes after translation is complete. The next three exons encode the extracellular 1, 2, and 3 domains, and the following downstream exon encodes the transmembrane (Tm) region. Finally, one or two 3-terminal exons encode the cytoplasmic domains (C). Like class I MHC genes, the class II genes are organized into a series of exons and introns mirroring the domain (a) α1 L α2 α3 Tm C 3′ L α1 α2 α3 Allelic Forms of MHC Genes Are Inherited in Linked Groups Called Haplotypes The genes that reside within the MHC region are highly polymorphic; that is, many alternative forms of each gene, or alleles, exist within the population. The individual genes of the MHC loci (class I, II, and III) lie so close together that their inheritance is linked. Crossover, or recombination between genes, is more likely when genes are far apart. For instance, the recombination frequency within the H-2 complex (i.e., the (b) C DNA 5′ structure of the  and  chains (Figure 8-9b). Both the  and the  genes encoding mouse and human class II MHC molecules have a leader exon, an 1 or 1 exon, an 2 or 2 exon, a transmembrane exon, and one or more cytoplasmic exons. L β1 L β1 (A)n β2 Tm+C C C (A) n mRNA Class II MHC molecule α3 S C β chain α2 S C 3′ α chain Class I MHC molecule Tm+C DNA 5′ Tm C C mRNA β2 S S S S COOH β1 S H2N H2N α1 β2 S α1 S S COOH S S COOH α2 H2N α chain β2 - microglobulin (A) n mRNA L α1 α2 Tm+C C α1 α2 DNA 5′ 3′ L FIGURE 8-9 Schematic diagram of (a) class I and (b) class II MHC genes, mRNA transcripts, and protein molecules. There is strong correspondence between exons and the domains in the gene products of MHC molecules. Note that the mRNA transcripts are spliced to remove the intron sequences. Each exon, with the excep- Tm+C C tion of the leader (L) exon, encodes a separate domain of an MHC molecule. The leader peptide is removed in a post-translational reaction before the molecule is expressed on the cell surface. The gene encoding 2-microglobulin is located on a different chromosome in both human and mouse. Tm  transmembrane; C  cytoplasmic. The Major Histocompatibility Complex and Antigen Presentation frequency of chromosome crossover events during meiosis, indicative of the distance between given genes) is only 0.5%. Thus, crossover occurs only once in every 200 meiotic cycles. For this reason, most individuals inherit all the alleles encoded by these genes as a set (known as linkage disequilibrium). This set of linked alleles is referred to as a haplotype. An individual inherits one haplotype from the mother and one haplotype from the father, or two sets of alleles. In outbred populations, such as humans, the offspring are generally heterozygous at the MHC locus, with different alleles contributed by each of the parents. If, however, mice are inbred, each H-2 locus becomes homozygous because the maternal and paternal haplotypes are identical, and all offspring begin to express identical MHC molecules. Certain mouse strains have been intentionally inbred in this manner and are employed as prototype strains. The MHC haplotype expressed by each of these strains is designated by an arbitrary italic superscript (e.g., H-2a, H-2b). These designations refer to the entire set of inherited H-2 alleles within a strain without having to list the specific allele at each locus individually (Table 8-2). Different inbred strains may share the same set of alleles, or MHC haplotype, with another strain (i.e., CBA, AKR, and C3H) but will differ in genes outside the H-2 complex. Detailed analysis of the H-2 complex in mice has been made possible by the development of congenic H-2 strains that differ only at the MHC locus. Inbred mouse strains are said to be syngeneic, or identical at all genetic loci. Two strains are congenic if they are genetically identical except at a single genetic region. Any phenotypic differences that can be detected between congenic strains is therefore related to the genetic region that distinguishes the two strains. Congenic strains that are identical with each other except at the TABLE 8-2 | CHAPTER 8 271 MHC can be produced by a series of crosses, backcrosses, and selections between two inbred strains that differ at the MHC. A frequently used congenic strain, designated B10.A, is derived from B10 mice (which is H-2b) genetically manipulated to possess the H-2a haplotype at the MHC locus. Recombination within the H-2 region of congenic mouse strains then allows the study of individual MHC genes and their products. Examples of these are included in the list in Table 8-2. For example, the B10.A (2R) strain has all the MHC genes from the a haplotype except for the D region, which is derived from the H-2b parent. MHC Molecules Are Codominantly Expressed The genes within the MHC locus exhibit a codominant form of expression, meaning that both maternal and paternal gene products (from both haplotypes) are expressed at the same time and in the same cells. Therefore, if two mice from inbred strains possessing different MHC haplotypes are mated, the F1 generation inherits both parental haplotypes and will express all these MHC alleles. For example, if an H-2b strain is crossed with an H-2k strain, then the F1 generation inherits both parental sets of alleles and is said to be H-2b/k (Figure 8-10a). Because such an F1 generation expresses the MHC proteins of both parental strains on its cells, it is said to be histocompatible with both parental strains. This means offspring are able to accept grafts from either parental source, each of which expresses MHC alleles viewed as “self ” (Figure 8-10b). However, neither of the inbred parental strains can accept a graft from its F1 offspring because half of the MHC molecules (those coming from the other parent) will be viewed as “nonself,” or foreign, and thus subject to recognition and rejection by the immune system. H-2 haplotypes of some mouse strains H-2 ALLELES Prototype strain Other strains with the same haplotype Haplotype K IA IE S D CBA AKR, C3H, B10.BR, C57BR k k k k k k DBA/2 BALB/c, NZB, SEA, YBR d d d d d d C57BL/10 (B10) C57BL/6, C57L, C3H.SW, LP, 129 b b b b b b A B10.A(2R)* B10.A(3R) B10A.(4R) A/He, A/Sn, A/Wy, B10.A a h2 i3 h4 k k b k k k b k k k k b d d d b d b d b A.SW B10.S, SJL s s s s s s t1 s k k k d q q q q q q A.TL DBA/1 STOLI, B10.Q, BDP *The R designates a recombinant haplotype, in this case between the H-2a and H-2b types. Gene contribution from the a strain is shown in yellow and from the b strain in red. FIGURE 8-10 Illustration of inheri- (a) Mating of inbred mouse strains with different MHC haplotypes Homologous chromosomes with MHC loci H-2b parent H-2k parent b/b b/b k/k k/k F1 progeny (H-2b/k) b/k b/k (b) Skin transplantation between inbred mouse strains with same or different MHC haplotypes Parental recipient Skin graft donor Progeny recipient b/b Parent b/k k/k Parent b/k b/k Progeny b/k b/b k/k b/b k/k b/b k/k (c) Inheritance of HLA haplotypes in a typical human family Parents A/B C/D Progeny A/C A/D B/R B/C B/D 272 tance of MHC haplotypes in inbred mouse strains. (a) The letters b/b designate a mouse homozygous for the H-2b MHC haplotype, k/k homozygous for the H-2k haplotype, and b/k a heterozygote. Because the MHC genes are closely linked and inherited as a set, the MHC haplotype of F1 progeny from the mating of two different inbred strains can be predicted easily. (b) Acceptance or rejection (X) of skin grafts is controlled by the MHC type of the inbred mice. The progeny of the cross between two inbred strains with different MHC haplotypes (H-2b and H-2k) will express both haplotypes (H-2b/k) and will accept grafts from either parent and from one another. However, neither parent strain will accept grafts from the offspring. (c) Inheritance of HLA haplotypes in a hypothetical human family. For ease, the human paternal HLA haplotypes are arbitrarily designated A and B, maternal C and D. Note that a new haplotype, R (recombination), can arise from rare recombination of a parental haplotype (maternal shown here). The Major Histocompatibility Complex and Antigen Presentation In an outbred population such as humans, each individual is generally heterozygous at each locus. The human HLA complex is highly polymorphic, and multiple alleles of each class I and class II gene exist. However, as with mice, the human MHC genes are closely linked and usually inherited as a haplotype. When the father and mother have different haplotypes, as in the example shown in Figure 8-10c, there is a one-in-four chance that siblings will inherit the same paternal and maternal haplotypes and therefore will be histocompatible (i.e., genetically identical at their MHC loci) with each other; none of the offspring will be fully histocompatible with the parents. Although the rate of recombination by crossover is low within the HLA complex, it still contributes significantly to the diversity of the loci in human populations. Genetic recombination can generate new allelic combinations, or haplotypes (see haplotype R in Figure 8-10c), and the high number of intervening generations since the appearance of humans as a species has allowed extensive recombination. As a result of recombination and other mechanisms for generating mutations, it is rare for any two unrelated individuals to have identical sets of HLA genes. This makes transplantation between individuals who are not identical twins quite challenging! To address this, clinicians begin by looking for family members who will be at least partially histocompatible with the patient, or they rely on donor databases to look for an MHC match. Even with partial matches, physicians still need to administer heavy doses of immunosuppressive drugs to inhibit the strong rejection responses that typically follow tissue transplantation due to differences in the MHC proteins (see Chapter 16). Class I and Class II Molecules Exhibit Diversity at Both the Individual and Species Levels As noted earlier, any particular MHC molecule can bind many different peptides (called promiscuity), which gives the host an advantage in responding to pathogens. Rather than relying on just one gene for this task, the MHC region has evolved to include multiple genetic loci encoding proteins with the same function. In humans, HLA-A, B, or C molecules can all present peptides to CD8 T cells and HLA-DP, DQ, or DR molecules can present to CD4 T cells. The MHC region is thus said to be polygenic because it contains multiple genes with the same function but with slightly different structures. Since the MHC alleles are also codominantly expressed, heterozygous individuals will express the gene products encoded by both alleles at each MHC gene locus. In a fully heterozygous individual this amounts to 6 unique classical class I molecules on each nucleated cell. An F1 mouse, for example, expresses the K, D, and L class I molecules from each parent (six different class I MHC molecules) on the surface of each of its nucleated cells (Figure 8-11). The expression of so many individual class I MHC molecules allows each cell to display a large number of different peptides in the peptide-binding grooves of its MHC molecules. | CHAPTER 8 Dk Class I molecules Dd Kd 273 Lk Kk Ld Maternal MHC Kk IAα k β k IEα k β k D k Lk I II I K d IAα d β d IEα d β d D d Ld IE αk β k IAα k β k Paternal MHC IE αd β d IAα d β d IE αk β d Class II molecules IE αd β k IAα k β d IAα d β k FIGURE 8-11 Diagram illustrating the various MHC molecules expressed on antigen-presenting cells of a heterozygous H-2k/d mouse. Both the maternal and paternal MHC genes are expressed (codominant expression). Because the class II molecules are heterodimers, new molecules containing one maternal-derived and one paternal-derived chain are also produced, increasing the diversity of MHC class II molecules on the cell surface. The 2-microglobulin component of class I molecules (pink) is encoded by a nonpolymorphic gene on a separate chromosome and may be derived from either parent. MHC class II molecules have even greater potential for diversity. Each of the classical class II MHC molecules is composed of two different polypeptide chains encoded by different loci. Therefore, a heterozygous individual can express - combinations that originate from the same chromosome (maternal only or paternal only) as well as class II molecules arising from unique chain pairing derived from separate chromosomes (new maternal-paternal - combinations). For example, an H-2k mouse expresses IAk and IEk class II molecules, whereas an H-2d mouse expresses IAd and IEd molecules. The F1 progeny resulting from crosses of these two strains express four parental class II molecules (identical to their parents) and also four new molecules that are mixtures from their parents, containing one parent’s  chain and the other parent’s  chain (as shown in Figure 8-11). Since the human MHC contains three classical class II genes (DP, DQ, and DR), a heterozygous individual expresses six class II molecules identical to the parents and six molecules containing new  and  chain combinations. The number of different class II molecules expressed by an individual can be further increased by the occasional presence of multiple -chain genes in mice and humans, and in humans by the presence of multiple -chain genes. The diversity generated by these new MHC molecules likely increases the number of different antigenic peptides that can be presented and is therefore advantageous to the organism in fighting infection. 274 PA R T I I I | Adaptive Immunity: Antigen Receptors and MHC The variety of peptides displayed by MHC molecules echoes the diversity of antigens bound by antibodies and T-cell receptors. This evolutionary pressure to diversify comes from the fact that both need to be able to interact with antigen fragments they have never before seen, or that may not yet have evolved. However, the strategy for generating diversity within MHC molecules and the antigen receptors on T and B cells is not the same. Antibodies and T-cell receptors are generated by several somatic processes, including gene rearrangement and the somatic mutation of rearranged genes (see Chapters 7 and 12). Thus, the generation of T- and B-cell receptors is dynamic, changing over time within an individual. By contrast, the MHC molecules expressed by an individual are fixed. However, promiscuity of antigen binding ensures that even “new” proteins are likely to contain some fragments that can associate with any given MHC molecule. TABLE 8-3 Genetic diversity of MHC loci in the human population MHC CLASS I HLA locus Number of allotypes (proteins) A 1448 B 1988 C 1119 E 3 F 4 G 16 MHC CLASS II HLA locus Number of allotypes (proteins) DMA 4 DMB 7 DOA 3 DOB 5 DPA1 17 DPB1 134 DQA1 47 DQB1 126 DRA 2 DRB1 860 DRB3 46 DRB4 8 DRB5 17 Source: Data obtained from http://hla.alleles.org, a Web site maintained by the HLA Informatics Group based at the Anthony Nolan Trust in the United Kingdom, with up-to-date information on the numbers of HLA alleles and proteins. Collectively, this builds in enormous flexibility within the host for responding to unexpected environmental changes that might arise in the future—an elegant evolutionary strategy. Countering this limitation on the range of peptides that can be presented by any one individual is the vast array of peptides that can be presented at the species level, thanks to the diversity of the MHC in any outbred population. The MHC possesses an extraordinarily large number of different alleles at each locus and is one of the most polymorphic genetic complexes known in higher vertebrates. These alleles differ in their DNA sequences from one individual to another by 5% to 10%. The number of amino acid differences between MHC alleles can be quite significant, with up to 20 amino acid residues contributing to the unique structural nature of each allele. Analysis of human HLA class I genes as of July 2012 reveals approximately 2013 A alleles, 2605 B alleles, and 1551 C alleles (Table 8-3 shows the number of protein products; not all alleles encode expressed proteins). In mice, the polymorphism is similarly enormous. The human class II genes are also highly polymorphic, and in some cases, different individuals can even inherit different numbers of genes. The number of HLA-DR -chain genes (DRB) may vary from 2 to 9 in different haplotypes, and over 1200 alleles of DRB genes alone have been reported. Interestingly, the DRA gene is highly conserved, with only seven different alleles and two proteins identified. Current estimates of actual polymorphism in the human MHC are likely on the low side because the earliest and most detailed data have primarily concentrated on populations of European descent. This enormous polymorphism results in a tremendous diversity of MHC molecules within a species. Using the numbers given above for the allelic forms of human HLA-A, -B, and -C, we can calculate a theoretical number of combinations that can exist, which is upward of 1.7 billion different class I haplotypes possible in the population. If class II loci are considered, the numbers are even more staggering, with over 1015 different class II combinations. Because each haplotype contains both class I and class II genes, multiplication of the numbers gives a total of more than 1.7  1024 possible combinations of these class I and II alleles within the entire human population! Some evidence suggests that a reduction in MHC polymorphism within a species may predispose that species to infectious disease (see Evolution Box 8-1). In one example, cheetahs and certain other wild cats, such as Florida panthers, that have been shown to be highly susceptible to viral disease also have very limited MHC polymorphism. It is postulated that the present cheetah population arose from a limited breeding population, or genetic bottleneck, causing a loss of MHC diversity. This increased susceptibility of cheetahs to various viruses may result from a reduction in the number of different MHC molecules available to the species as a whole and a corresponding limitation on the range of processed antigens with which these MHC molecules can interact. As a corollary, this suggests that the high level of MHC polymorphism that has been observed in many outbred species, including humans, may provide a survival advantage by supplying a broad range of MHC molecules and thus a broad range of presentable antigens. The Major Histocompatibility Complex and Antigen Presentation | CHAPTER 8 275 BOX 8-1 EVOLUTION The Sweet Smell of Diversity As early as the mid-1970s, mate choice in mice was shown to be influenced by genes at the MHC (H-2) locus. This was followed by investigations in other rodents, fish, and birds, all with similar conclusions. Thanks to results from several studies conducted in the last 15 years, it looks like humans could be added to this list. However, questions remain as to the precise evolutionary pressures, the mechanism, and the magnitude of this effect, among others, in influencing mate choice in humans. In terms of evolutionary pressure, local pathogens play a significant role in maintaining MHC diversity in the population and in selecting for specific alleles. As we now appreciate, this is because the MHC influences immune responsiveness. Due to its role in selecting the peptide fragments that will be presented, the inheritance of specific alleles at particular loci can predispose individuals to either enhanced susceptibility or resistance to specific infectious agents and immune disorders (see Clinical Focus Box 8-2). Over long time periods, endemic local pathogens exert evolutionary pressures that drive higher- or lower-than-expected rates of certain MHC alleles in a population, as well as overrepresentation of other resistance-associated genes. The degree of diversity at the MHC locus clearly influences susceptibility to disease in populations; witness the enhanced viral susceptibility seen in cheetahs (see page 274) and some devastating human stories. For instance, the European introduction of the smallpox virus to New World populations is credited with wiping out large Native American groups. This may be due to a lack of past evolutionary pressure for conservation of resistance-associated MHC alleles, which would therefore be rare or nonexistent in this population, as well as to a lack of any individuals with immunity from prior infection. But how does an individual evaluate how well a potential mate will contribute “new” MHC alleles to one’s offspring, leading to greater diversity and the potential for enhanced fitness? The primary candidate is odor: the MHC is known to influence odor in many vertebrate species. For example, the urine of mice from distinct MHC congenic lines can be distinguished by both humans and rodents. Mice show a distinct preference for mating with animals that carry MHC alleles that are dissimilar to their own. In terms of maximizing the range of peptides that can be presented, this makes clear evolutionary sense, as increased diversity at the MHC should increase the number of different pathogenic peptides that can be “seen” by the immune system, increasing the likelihood of effective anti-pathogen responses. To back this up, the advantage of overall MHC diversity has been shown experimentally in mice, where most simulated epidemic experiments have found a survival advantage for H-2 heterozygous animals over their homozygous counterparts. In humans, research in HIV-infected individuals has shown that extended survival and a slower progression to AIDS are correlated with full heterozygosity at the HLA class I locus, as well as absence of certain AIDS-associated HLA-B and -C alleles. This specific link to class I is not surprising in light of the key role of CD8 T cells in combating viral infections. Human studies of attraction and mating also point to preferences for individuals with MHC dissimilar alleles. In one key study involving what is commonly known as the “sweaty T-shirt test,” college-age volunteers were asked to rate their preference or sexual attraction to the odor of T-shirts worn by individuals of the opposite sex. In general, both males and females preferred the odor of T-shirts worn by individuals with dissimilar HLA types. The one key exception was seen in women who were concurrently using an estrogen-based birth control pill; they instead showed a preference for the odors of MHC-similar individuals, suggesting that this hormone not only interferes with this response but that it potentially shifts the outcome. But how? you ask. Soluble forms of MHC molecules have been found in many bodily fluids, including urine, saliva, sweat, and plasma. However, these molecules are unlikely to be small or volatile enough to account for direct olfactory detection. Other hypotheses for how MHC might influence odor include via olfactory recognition of natural ligands or specific volatile peptides carried by MHC molecules, or by MHC-driven differences in natural flora, which are also known to impact body odor. The recent finding that some polymorphic olfactory receptor genes are closely linked to the MHC may also help explain this apparent association with mating preference. To date, the jury is still out on which mechanism(s) likely account(s) for MHCspecific odor differences. As one might imagine, large methodological challenges are inherent in asking questions related to mate choice in humans, where differing levels of outbreeding as well as social and cultural factors influence the outcome. Nonetheless, it appears that like most primate species tested so far, we humans may also have the capacity to use smell as an evolutionary strategy for promoting a robust immune response in our offspring. J. Havlicek, and S. C. Roberts. 2009. MHC-correlated mate choice in humans: A review. Psychoneuroendocrinology 34:497–512. C. Wedekind et al. 1995. MHC-dependent mate preferences in humans. Proceedings. Biological Science 260:245–249. C. Wedekind et al. 1997. Body odor preferences in men and women: Do they aim for specific MHC combinations or simply heterozygosity? Proceedings. Biological Science 264:1471– 1479. 276 PA R T I I I | Adaptive Immunity: Antigen Receptors and MHC Although some individuals within a species may not be able to develop an immune response to a given pathogen and therefore will be susceptible to infection, extreme polymorphism ensures that at least some members of a species will be resistant to that disease. In this way, MHC diversity at the population level may protect the species as a whole from extinction via a wide range of infectious diseases. MHC Polymorphism Has Functional Relevance Although the sequence divergence among alleles of the MHC within a species is very high, this variation is not randomly distributed along the entire polypeptide chain. Instead, polymorphism in the MHC is clustered in short stretches, largely within the membrane-distal 1 and 2 domains of class I molecules (Figure 8-12a). Similar patterns of diversity are observed in the 1 and 1 domains of class II molecules. (a) α2 α3 Variability α1 Structural comparisons have located the polymorphic residues within the three-dimensional structure of the membranedistal domains in class I and class II MHC molecules and have related allelic differences to functional differences (Figure 8-12b). For example, of 17 amino acids previously shown to display significant polymorphism among HLA-A molecules, 15 were shown by x-ray crystallographic analysis to be in the peptide-binding groove of this molecule. The location of so many polymorphic amino acids within the binding site for processed antigen strongly suggests that allelic differences contribute to the observed differences in the ability of MHC molecules to interact with a given peptide ligand. Polymorphisms that lie outside these regions and might affect basic domain folding are rare. This clustering of polymorphisms around regions that make contact with antigen also suggest possible reasons why certain MHC genes or haplotypes can become associated with certain diseases (see Clinical Focus Box 8-2). 20 40 60 80 100 120 140 160 Residue number 180 200 220 240 260 (b) 12 45 62 63 66 70 74 95 9 97 116 114 156 N 107 105 FIGURE 8-12 Variability in the amino acid sequences of allelic class I HLA molecules. (a) In the external domains, most of the variable residues are in the membrane-distal 1 and 2 domains. (b) Location of polymorphic amino acid residues (red) in the 1/2 domain of a human class I MHC molecule. [Part (a) adapted from R. Sodoyer et al., 1984, EMBO Journal 3:879, reprinted by permission of Oxford University Press; part (b) adapted, with permission, from P. Parham, 1989, Nature 342:617, © 1989 Macmillan Magazines Limited.] The Major Histocompatibility Complex and Antigen Presentation | CHAPTER 8 277 BOX 8-2 CLINICAL FOCUS MHC Alleles and Susceptibility to Certain Diseases Some HLA alleles occur at a much higher frequency in people suffering from certain diseases than in the general population. The diseases associated with particular MHC alleles include autoimmune disorders, certain viral diseases, disorders of the complement system, some neurologic disorders, and several different allergies. In humans, the association between an HLA allele and a given disease may be quantified by determining the frequency of that allele expressed by individuals afflicted with the disease, then comparing these data with the frequency of the same allele in the general population. Such a comparison allows calculation of an individual’s relative risk (RR). RR  frequency of disease in the allele group frequency of disease in the allele group (Eq. 8.1) A RR value of 1 means that the HLA allele is expressed with the same frequency in disease-afflicted and general populations, indicating that this allele confers no increased risk for the disease. A RR value substantially above 1 indicates an association between the HLA allele and the disease. For example, individuals with the HLA-B27 allele are 90 times more likely (RR  90) to develop the autoimmune disease ankylosing spondylitis, an inflammatory disease of vertebral joints characterized by destruction of cartilage, than are individuals who lack this HLA-B allele. Other disease associations with significantly high RR include HLA-DQB1 and narcolepsy (RR  130) and HLA-DQ2 with celiac disease (RR50), an allergy to gluten. A few HLA alleles have also been linked to relative protection from disease or clinical progres- sion. This is seen in the case of individuals inheriting HLA-B57, which is associated with greater viral control and a slower progression to AIDS in HIV-infected individuals. When the associations between MHC alleles and disease are weak, reflected by low RR values, it is possible that multiple genes influence susceptibility, of which only one lies within the MHC locus. The genetic origins of several autoimmune diseases, such as multiple sclerosis (associated with DR2; RR = 5) and rheumatoid arthritis (associated with DR4; RR = 10) have been studied in depth. The observation that these diseases are not inherited by simple Mendelian segregation of MHC alleles can be seen clearly in identical twins, where both inherit the same MHC risk factor, but frequently only one develops the disease. This finding suggests that multiple genetic factors plus one or more environmental factors are at play in development of this disease. As we highlight in Chapter 16, this combined role for genes and the environment in the development of autoimmunity is not uncommon. The existence of an association between an MHC allele and a disease should not be interpreted to imply that the expression of the allele has caused the disease. The relationship between MHC alleles and development of disease is more complex, partly thanks to the genetic phenomenon of linkage disequilibrium. The fact that some of the class I MHC alleles are in linkage disequilibrium with the class II and class III alleles can make their contribution to disease susceptibility appear more pronounced than it actually is. If, for example, DR4 contrib- The preceding discussion points to additional parallels between MHC molecules and lymphocyte antigen receptors. The somatic hypermutations seen in B-cell receptor genes is also not randomly arrayed within the molecule, but instead is clustered in the regions most likely to interact directly with peptide (see Chapter 12), providing yet another example of how the immune system has solved a similar functional dilemma using a very different strategy. utes to risk of a disease and it also occurs frequently in combination with HLA-A3 because of linkage disequilibrium, then A3 would incorrectly appear to be associated with the disease. Improved genomic mapping techniques make it possible to analyze the fine linkage between genes within the MHC and various diseases more fully and to assess the contributions from other loci. In the case of ankylosing spondylitis, for example, it has been suggested that alleles of TNF-␣ and Lymphotoxin-␣ may produce protein variants that are involved in destruction of cartilage, and these alleles happen to be linked to certain HLA-B alleles. In the case of HLAB57 and AIDS progression, the allele has been linked more directly to disease. It is believed that this class I allele is particularly efficient at presenting important components of the virus to circulating Tc cells, leading to increased destruction of virally infected cells and delayed disease progression. Other hypotheses have also been offered to account for a direct role of particular MHC alleles in disease susceptibility. In some rare cases, certain allelic forms of MHC genes can encode molecules that are recognized as receptors by viruses or bacterial toxins, leading to increased susceptibility in the individuals who inherit these alleles. As will be explored further in Chapters 15 through 19, in many cases complex interactions between multiple genes, frequently including the MHC, and particular environmental factors are required to create a bias toward the development of certain diseases. The Role of the MHC and Expression Patterns As we have just discussed, several genetic features help ensure a diversity of MHC molecules in outbred populations, including polygeny, polymorphism, and codominant expression. All this attention paid to maximizing the number of different 278 PA R T I I I | Adaptive Immunity: Antigen Receptors and MHC binding grooves suggests that variety within the MHC plays an important role in survival (see Clinical Focus Box 8-2). In fact, in addition to fighting infection, MHC expression throughout the body plays a key role in maintaining homeostasis and health even when no foreign antigen is present. Although the presentation of MHC molecules complexed with foreign antigen to T cells garners much attention (and space in this book!), most MHC molecules spend their lives presenting other things, and often to other cells. There are several reasons why an MHC molecule on the surface of a cell is important. In general, these include the following: • To display self class I to demonstrate that the cell is healthy • To display foreign peptide in class I to show that the cell is infected and to engage with TC cells • To display a self-peptide in class I and II to test developing T cells for autoreactivity (primary lymphoid organs) • To display a self-peptide in class I and II to maintain tolerance to self-proteins (secondary lymphoid organs) • To display a foreign peptide in class II to show the body is infected and activate TH cells It is worth noting that although some of these instances lead to immune activation, some do not. It is also important to note that the cell type, tissue location, and timing of expression vary for each of these situations. For instance, developing T cells first encounter MHC class I and II molecules presenting self peptides in the thymus, where these signals are designed to inhibit the ability of T cells to attack selfstructures later (see Chapter 9). Others occur only after certain types of exposure, such as during an immune response to extracellular pathogens or when cells in the body are infected with viral invaders, and is designed to activate T cells to act against the pathogen (see Chapter 11). To help set the stage for this discussion, we turn now to the where, when, and why of MHC expression. This is followed by a description of the how, detailing the pathways that lead to peptide placement in the binding groove of each class of molecule. As we will see in greater detail in the following sections, the type of MHC molecule(s) expressed by a given cell is linked to the role of the cell and the function of each class of molecule. MHC Molecules Present Both Intracellular and Extracellular Antigens In very general terms, the job of MHC class I molecules is to collect and present antigens that come from intracellular locations. This is a form of ongoing surveillance of the internal happenings of the cell—in essence, a window for displaying on the surface of the cell snippets of what is occurring inside. Basically, all cells in the body need this form of check and balance, and this shows up in the fairly ubiquitous nature of MHC class I expression in the body (there are a few notable exceptions, which we will touch on later). Often, nothing other than normal cellular processes are occurring in the cytosol, and in these instances our cells present self-peptides in the groove of MHC class I molecules. The expresssion of self-MHC class I (with self peptides) signals that a cell is healthy; absence of self-MHC class I (as can occur in virus-infected and tumor cells) targets that cell for killing by NK cells (see Chapter 13). When foreign proteins are present in the cytosol and begin to appear in the groove of MHC class I on the cell surface, this alerts CD8 T cells to the presence of this unwelcome visitor, targeting the cell for destruction. In this case, the cell is called a target cell because it becomes a target for lysis by cytotoxic T cells. Conversely, MHC class II molecules primarily display peptides that have come from the extracellular spaces of the body. Since this sampling of extracellular contents is not a form of policing that all cells need to perform, only specialized leukocytes posses this ability. These cells are collectively referred to as antigen-presenting cells (APCs), because their job is to present extracellular antigen to T cells, charging them with the ultimate job of coordinating the elimination of this extracellular invader. While all cells in the body express MHC class I proteins and can present peptides from foreign intracellular antigens, the term APC is usually reserved for MHC class II-expressing cells. MHC Class I Expression Is Found Throughout the Body In general, classical class I MHC molecules are expressed constitutively on almost all nucleated cells of the body. However, the level of expression differs among different cell types, with the highest levels of class I molecules found on the surface of lymphocytes. These molecules may constitute approximately 1% of the total plasma membrane proteins, or some 5  105 MHC class I molecules per lymphocyte. In contrast, some other cells, such as fibroblasts, muscle cells, liver hepatocytes, and some neural cells, express very low to undetectable levels of class I MHC molecules. This low-level expression on liver cells may contribute to the relative success of liver transplants by reducing the likelihood of graft rejection by Tc cells of the recipient. A few cell types (e.g., subsets of neurons and sperm cells at certain stages of differentiation) appear to lack class I MHC molecules altogether. Nucleated cells without class I expression are, however, very rare. Non-nucleated cells, such as red blood cells in mammals, do not generally express any MHC molecules. In normal, healthy cells, class I molecules on the surface of the cell will display self-peptides resulting from normal turnover of self-proteins inside the cell. In cells infected by a virus, viral peptides as well as self-peptides will be displayed. Therefore, a single virus-infected cell can be envisioned as having various class I molecules on its membrane, some displaying a subset of viral peptides derived from the viral proteins being manufactured within. Because of individual allelic differences in the peptide-binding grooves of the class I MHC molecules, different individuals within a species will have the ability to bind and present different sets of viral The Major Histocompatibility Complex and Antigen Presentation peptides. In addition to virally infected cells, altered self cells such as cancer cells, aging body cells, or cells from an allogeneic graft (i.e., from a genetically-different individual), also can serve as target cells due to their expression of foreign MHC proteins and can be lysed by TC cells. The importance of constitutive expression of class I is highlighted by the response of natural killer (NK) cells to somatic cells that lack MHC class I, as can occur during some viral infections. NK cells can kill a cell that has stopped expressing MHC class I on the surface, presumably because this suggests that the cell is no longer healthy or has been altered by the presence of an intracellular invader. Expression of MHC Class II Molecules Is Primarily Restricted to Antigen-Presenting Cells MHC class II molecules are found on a much more restricted set of cells than class I, and sometimes only after an inducing event. As mentioned above, cells that display peptides associated with class II MHC molecules to CD4 TH cells are called antigen-presenting cells (APCs), and these cells are primarily certain types of leukocytes. APCs are specialized for their ability to alert the immune system to the presence of an invader and drive the activation of T cell responses. Among the various APCs, marked differences in the level of MHC class II expression have been observed. In some cases, class II expression depends on the cell’s differentiation stage or level of activation (such as in macrophages; see below). APC activation usually occurs following interaction with a pathogen and/or via cytokine signaling, which then induces significant increases in MHC class II expression. A variety of cells can function as bona fide APCs. Their distinguishing feature is their ability to express class II MHC molecules and to deliver a costimulatory, or second activating signal, to T cells. Three cell types are known to have these characteristics and are thus often referred to as professional antigen-presenting cells (pAPCs): dendritic cells, macrophages, and B lymphocytes. These cells differ from one another in their mechanisms of antigen uptake, in whether they constitutively express class II MHC molecules, and in their costimulatory activity, as follows: • Dendritic cells are generally viewed as the most effective of the APCs. Because these cells constitutively express a TABLE 8-4 Antigen-presenting cells Professional antigenpresenting cells Nonprofessional antigenpresenting cells Dendritic cells (several types) Fibroblasts (skin) Thymic epithelial cells Macrophages Glial cells (brain) Thyroid epithelial cells B cells Pancreatic beta cells Vascular endothelial cells | CHAPTER 8 279 high levels of class II MHC molecules and have inherent costimulatory activity, they can activate naïve TH cells. • Macrophages must be activated (e.g., via TLR signaling) before they express class II MHC molecules or costimulatory membrane molecules such as CD80/86. • B cells constitutively express class II MHC molecules and posses antigen-specific surface receptors, making them particularly efficient at capturing and presenting their cognate antigen. However, they must be activated by, for example, antigen, cytokines, or pathogen-associated molecular patterns (PAMPs), before they express the costimulatory molecules required for activating naïve TH cells. Several other cell types, classified as nonprofessional APCs, can be induced to express class II MHC molecules and/or a costimulatory signal under certain conditions (Table 8-4). These cells can be deputized for professional antigen presentation for short periods and in particular situations, such as during a sustained inflammatory response. MHC Expression Can Change with Changing Conditions As noted above, MHC class I is constitutively expressed by most cells in the body, whereas class II is only expressed under certain conditions and in a very limited number of cell types. The different roles of the two molecules help explain this, and suggest that in certain instances specific changes in MHC expression may prove advantageous. The MHC locus can respond to both positive and negative regulatory pressures. For instance, MHC class I production can be disrupted or depressed by some pathogens. MHC class II expression on APCs is already quite variable, and the microenvironment surrounding an APC can further modulate expression, usually enhancing the expression of these molecules. APCs in particular are conditioned to respond to local cues leading to their activation and heightened MHC expression, enabling them to arm other cells in the body for battle. As one might imagine, this activation and arming of APCs must be carefully regulated, lest these cells orchestrate unwanted aggressive maneuvers against self components or benign companions, as occurs in clinical conditions such as autoimmunity or allergy, respectively. The mechanisms driving these changes in expression are described below. Genetic Regulatory Components The presence of internal or external triggers, such as intracellular invaders or cytokines (see below), can induce a signal transduction cascade that leads to changes in MHC gene expression. Research aimed at understanding the mechanism of control of MHC expression has been advanced by the now complete sequence of the mouse genome. Both class I and class II MHC genes are flanked by 5 promoter sequences that bind sequence-specific transcription factors. The promoter 280 PA R T I I I | Adaptive Immunity: Antigen Receptors and MHC motifs and the transcription factors that bind to these motifs have been identified for a number of MHC genes, with examples of regulation mediated by both positive and negative elements. For example, a class II MHC transcriptional activator called CIITA (also known as class II, major histocompatibility complex, transactivator) and another transcription factor called RFX, have both been shown to activate the promoter of class II MHC genes. Defects in these transcription factors cause one form of bare lymphocyte syndrome. Patients with this disorder lack class II MHC molecules on their cells and suffer from severe immunodeficiency, highlighting the central role of class II molecules in T-cell maturation and activation. Viral Interference One clear example of negative regulation of MHC comes from viruses that interfere with MHC class I expression and thus avoid easy detection by CD8 T cells. These viruses include human cytomegalovirus (CMV), hepatitis B virus (HBV), and adenovirus 12 (Ad12). In some cases, reduced expression of class I MHC molecules is due to decreased levels of a component needed for peptide transport or MHC class I assembly rather than decreased transcription. For example, in the case of cytomegalovirus infection, a viral protein binds to 2-microglobulin, preventing assembly of class I MHC molecules and their transport to the plasma membrane. Adenovirus 12 infection causes a pronounced decrease in transcription of the transporter genes (TAP1 and TAP2). As described in the following section on antigen processing, the TAP gene products play an important role in peptide transport from the cytoplasm into the rough endoplasmic reticulum (RER). Blocking of TAP gene expression inhibits peptide transport; as a result, class I MHC molecules cannot assemble with 2-microglobulin or be transported to the cell membrane. These observations are especially important because decreased expression of class I MHC molecules, by whatever mechanism, is likely to help viruses evade the immune response. Lower levels of class I decrease the likelihood that virus-infected cells can display viral peptide complexes on their surface and become targets for CTL-mediated destruction. Cytokine-Mediated Signaling The expression of MHC molecules is externally regulated by various cytokines. Leading among these are the interferons (, , and ) and the tumor necrosis factors (TNF- and Lymphotoxin-), each of which have been shown to increase expression of class I MHC molecules on cells. Typically, phagocytic cells that are involved in innate responses, or locally infected cells, are the first to produce these MHCregulating cytokines. In particular, IFN- (produced by a cell following viral or bacterial infection) and TNF- (secreted by APCs after activation) are frequently the first cytokines to kick off an MHC class I up-regulation event. In the later stages, interferon gamma (IFN- ), secreted by activated TH cells as well as other cell types, also contributes to increased MHC expression. Binding of these cytokines to their respective receptors induces intracellular signaling cascades that activate transcription factors and alter expression patterns. These factors bind to their target promoter sequences and coordinate increased transcription of the genes encoding the class I  chain, 2-microglobulin, and other proteins involved in antigen processing and presentation. IFN- has been shown to induce expression of the class II transcriptional activator (CIITA), thereby indirectly increasing expression of class II MHC molecules on a variety of cells, including non-APCs (e.g., skin keratinocytes, intestinal epithelial cells, vascular endothelium, placental cells, and pancreatic beta cells). Other cytokines influence MHC expression only in certain cell types. For example, IL-4 increases expression of class II molecules in resting B cells, turning them into more efficient APCs. Conversely, expression of class II molecules by B cells is down-regulated by IFN- . Corticosteroids and prostaglandins can also decrease expression of class II MHC molecules. These naturally occurring, membrane-permeable compounds bind to intracellular receptors and are some of the most potent suppressors of adaptive immunity, primarily based on this ability to inhibit MHC expression. This property is exploited in many clinical settings, where these compounds are used as treatments to suppress overly zealous immune events, such as in allergic responses or during transplant rejection (see Chapters 15 and 16). Class II MHC Alleles Play a Critical Role in Immune Responsiveness MHC haplotype plays a strong role in the outcome of an immune response, as these alleles determine which fragments of protein will be presented. Recall that class II MHC molecules present foreign antigen to CD4 TH cells, which go on to activate B cells to produce serum antibodies. Early studies by Benacerraf in which guinea pigs were immunized with simple synthetic antigens were the first to show that the ability of an animal to mount an immune response, as measured by the production of serum antibodies, is determined by its MHC haplotype. Later experiments by H. McDevitt, M. Sela, and their colleagues used congenic (matched at the MHC locus but not elsewhere) and recombinant congenic mouse strains to specifically map the control of immune responsiveness to class II MHC genes. In early reports, the genes responsible for this phenotype were designated Ir or immune response genes; retaining the initial I, mouse class II products are now called IA and IE. We now know that the dependence of immune responsiveness on the genes within the class II MHC reflects the central role of these molecules in determining which fragments of foreign proteins are presented as antigen to TH cells. Two explanations have been proposed to account for this variability in immune responsiveness observed among different haplotypes. According to the determinant-selection model, different class II MHC molecules differ in their ability to bind particular processed antigens. In the end, some peptides may be more crucial to eliminating the pathogen than others. A separate hypothesis, termed holes-in-the-repertoire model, postulates that T cells bearing receptors that recognize | The Major Histocompatibility Complex and Antigen Presentation certain foreign antigens which happen to closely resemble self-antigens may be eliminated during T-cell development, leaving the organism without these cells/receptors for future responses to foreign molecules. Since the T-cell response to an antigen involves a trimolecular complex of the T-cell receptor, an antigenic peptide, and an MHC molecule (discussed in detail in Chapter 11), both models appear correct. That is, the absence of an MHC molecule that can bind and present a particular peptide, or the absence of T-cell receptors that can recognize a given MHC-peptide molecule complex, both result in decreased immune responsiveness to a given foreign substance, and can account for the observed relationship between MHC haplotype and the ability to respond to particular exogenous antigens. CHAPTER 8 Antigen Strain 2 or 13 or (2 × 13) F1 Strain 2 or 13 or (2 × 13) F1 7 days Lymph node cells (source of T cells) Peritoneal exudate cells (source of macrophages) Adherence column (retains macrophages) Adherent cells T Cells Are Restricted to Recognizing Peptides Presented in the Context of Self-MHC Alleles In the 1970s a series of experiments were carried out to further explore the relationship between the MHC and immune response. These investigations contributed two very crucial discoveries: (1) that both CD4 and CD8 T cells can recognize antigen only when it is presented in the groove of an MHC molecule, and (2) that the MHC haplotype of the APC and the T cell must match. This happens naturally in the host, where T cells develop alongside host APCs, both expressing only that individual’s MHC molecules (see Chapter 9). This constraint is referred to as self-MHC restriction, which refers to the dual specificity of T cells for self MHC as well as for foreign antigen. A. Rosenthal and E. Shevach showed that antigen-specific proliferation of TH cells occurs only in response to antigen presented by macrophages of the same MHC haplotype as the T cells recognizing the antigen. In their experimental system, guinea pig macrophages from strain 2 were initially incubated with an antigen (Figure 8-13). After the “antigenpulsed” macrophages had processed the antigen and presented it on their surface, they were mixed with T cells from the same strain (strain 2), a different strain (strain 13), or their F1 progeny (2 X 13), and the magnitude of T-cell proliferation in response to the antigen-pulsed macrophages was measured. The results of these experiments showed that strain-2 antigen-pulsed macrophages activated T cells from strain-2 and F1 mice but not T cells from strain-13 animals. Similarly, strain-13 antigen-pulsed macrophages activated strain-13 and F1 T cells but not strain-2 T cells. Subsequently, congenic and recombinant congenic strains of mice, which differed from each other only in selected regions of the MHC were used as the source of macrophages and T cells. These experiments confirmed that the CD4 TH cell is activated and proliferates only in the presence of antigenpulsed macrophages that share class II MHC alleles. Thus, antigen recognition by the CD4 TH cell is said to be class II MHC restricted. In 1974, Zinkernagel and Doherty similarly demonstrated the self-MHC class I restriction of CD8 T cells (see Classic Experiment Box 8-3). Over two decades later, the pair was awarded the Nobel Prize in Medicine for their 281 Peritoneal macrophages Antigen Antigen-primed T cells Antigen-pulsed macrophages Measure T-cell proliferation Antigen-primed T cell Strain 2 Strain 13 (2 × 13) F1 Antigen-pulsed macrophages Strain 2 Strain 13 (2 × 13) F1 + − + − + + + + + FIGURE 8-13 Experimental demonstration of self-MHC restriction of TH cells. Peritoneal exudate cells from strain 2, strain 13, or (2 X 13) F1 guinea pigs were incubated in plastic petri dishes, allowing enrichment of macrophages, which are adherent cells. The peritoneal macrophages were then incubated with antigen. These “antigen-pulsed” macrophages were incubated in vitro with T cells from strain 2, strain 13, or (2 X 13) F1 guinea pigs, and the degree of T-cell proliferation was assessed ( vs ). The results indicated that TH cells could proliferate only in response to antigen presented by macrophages that shared MHC alleles. [Adapted from A. Rosenthal and E. Shevach, 1974, Journal of Experimental Medicine 138:1194, by copyright permission of the Rockefeller University Press.] seminal studies (see Table 1-2 for a list of Nobel Prizes related to Immunology). As we will see in Chapter 9, selfMHC restriction is the result of T-cell development in the thymus, where only T cells that recognize self-MHC are selected for survival. 282 PA R T I I I | Adaptive Immunity: Antigen Receptors and MHC BOX 8-3 CLASSIC EXPERIMENT Demonstration of the Self-MHC Restriction of CD8 T Cells In 1974, R. M. Zinkernagel and P. C. Doherty demonstrated that CD8 T cells are restricted to recognizing antigen in the context of self-MHC molecules. For this seminal study, they took advantage of recently derived inbred mice strains, a virus that causes widespread neurological damage in infected animals, and new assays that allowed quantification of cytotoxic T cell responses. In their experiments, mice were first immunized with lymphocytic choriomeningitis virus (LCMV), a pathogen that results in central nervous system inflammation and neurological damage. Several days later, the animals’ spleen cells, which included TC cells specific for the virus, were isolated and incubated with LCMV-infected target cells of the same or different haplotype (Figure 1). The assay relied on measuring the release of a radioisotope (51Cr) from labeled target cells (called a chromium release assay; see Figure 13-18). They found that the TC cells killed syngeneic virus-infected target cells (or cells with matched MHC LCMV H–2k Spleen cells (containing Tc cells) 51Cr H–2k target cells –51Cr release (no lysis) H–2k LCMV-infected target cells H–2b LCMV-infected target cells +51Cr release (lysis) –51Cr release (no lysis) FIGURE 1 Classic experiment of Zinkernagel and Doherty demonstrating that antigen recognition by TC cells exhibits MHC restriction. H-2k mice were primed with the lymphocytic choriomeningitis virus (LCMV) to induce cytotoxic T lymphocytes (CTLs) specific for the virus. Spleen cells from this LCMV-primed mouse were then added to target cells of different H-2 haplotypes that were intracellularly labeled with 51Cr (black dots) and either infected or not with the LCMV. CTL-mediated killing of the target cells, as measured by the release of 51Cr into the culture supernatant, occurred only if the target cells were infected with LCMV and had the same MHC haplotype as the CTLs. [Adapted from P. C. Doherty and R. M. Zinkernagel, 1975, Journal of Experimental Medicine 141:502.] alleles) but not uninfected cells or infected cells from a donor that shared no MHC alleles with these cytotoxic cells. Later studies with congenic and recombinant congenic strains showed that the TC cell and the virus-infected target cell must specifically share class I molecules encoded by the K or D regions of the MHC. Thus, antigen recognition by CD8 cytotoxic T cells is class I MHC restricted. In 1996, Doherty and Zinkernagel were awarded the Nobel Prize in Medicine for their major contribution to understanding the role of the MHC in cellmediated immunity. Before this discovery, histocompatibility antigens were mainly blamed for transplantation rejection, but their importance in the everyday process of cellular immunity was not appreciated. Zinkernagel and Doherty’s discovery is all the more revolutionary when one takes into consideration that little was known about cellular immunity and no T-cell receptor had yet been discovered! This milestone in immunologic understanding set the stage for the development of two key models of how T cells respond to foreign antigen: altered self and dual recognition. The altered self hypothesis posited that histocompatibility molecules that associate with foreign particles, such as viruses, may appear to be altered forms of self-proteins. The dual recognition model proposed that T cells must be capable of simultaneous recognition of both the foreign substance and these selfhistocompatibility molecules. We now appreciate the validity of both of these models, and the role of MHC restriction in the immune response to both microorganisms and allogeneic transplants. Doherty, P. C., and R. M. Zinkernagel. 1975. H-2 compatibility is required for T-cell mediated lysis of target cells infected with lymphocytic choriomeningitis virus. Journal of Experimental Medicine 141:502. The Major Histocompatibility Complex and Antigen Presentation Processing of Antigen Is Required for Recognition by T Cells In the 1980s, K. Ziegler and E. R. Unanue showed that an intracellular processing step by APCs was required to activate T cells. These researchers observed that TH-cell activation by bacterial protein antigens was prevented by treating the APCs with paraformaldehyde prior to antigen exposure (in essence, killing the cells and immobilizing the antigens in the membrane; see Figure 8-14a). However, if the APCs were first allowed to ingest the antigen and were fixed with paraformaldehyde 1 to 3 hours later, TH-cell activation still occurred (Figure 8-14b). During that interval of 1 to 3 hours, the APCs had taken up the antigen, processed it into peptide fragments, and displayed these fragments on the cell membrane in a form capable of activating T cells. Subsequent experiments by R. P. Shimonkevitz showed that internalization and processing could be bypassed if APCs were exposed to already digested peptide fragments instead of the native antigen (Figure 8-14c). In these experi- | CHAPTER 8 283 ments, APCs were treated with glutaraldehyde (this chemical, like paraformaldehyde, fixes the cell, rendering it metabolically inactive) and then incubated with native ovalbumin or with ovalbumin that had been subjected to partial enzymatic digestion. The digested ovalbumin was able to interact with the glutaraldehyde-fixed APCs, thereby activating ovalbumin-specific TH cells. However, the native ovalbumin failed to do so. These results suggest that antigen processing requires the digestion of the protein into peptides that can be recognized by the ovalbumin-specific TH cells. At about the same time, several investigators, including W. Gerhard, A. Townsend, and their colleagues, began to identify the proteins of influenza virus that were recognized by TC cells. Contrary to their expectations, they found that internal proteins of the virus, such as polymerase and nucleocapsid proteins, were often recognized by TC cells better than the more exposed envelope proteins found on the surface of the virus. Moreover, Townsend’s work revealed that TC cells recognized short linear peptide sequences of influenza proteins. EXPERIMENTAL CONDITIONS T-CELL ACTIVATION (a) Antigen – 1h APC APC Fixation (b) Antigen Fixation + 1h TH cell APC APC Fixation APC (c) + Antigen peptides APC FIGURE 8-14 Experimental demonstration that antigen processing is necessary for TH-cell activation. (a) When antigenpresenting cells (APCs) are fixed before exposure to antigen, they are unable to activate TH cells. (b) In contrast, APCs fixed at least 1 hour after antigen exposure can activate TH cells. (This simplified figure TH cell APC does not show costimulatory molecules needed for T-cell activation.) (c) When APCs are fixed before antigen exposure and incubated with peptide digests of the antigen (rather than native antigen), they also can activate TH cells. TH-cell activation is determined by measuring a specific TH-cell response (e.g., cytokine secretion). 284 PA R T I I I | Adaptive Immunity: Antigen Receptors and MHC In fact, when noninfected target cells were incubated in vitro with synthetic peptides corresponding to only short sequences of internal influenza proteins, these cells could be recognized by TC cells and subsequently lysed just as well as target cells that had been infected with whole, live influenza virus. These findings, along with those presented below, suggest that antigen processing is a metabolic process that digests proteins into peptides, which can then be displayed on the cell surface together with a class I or class II MHC molecule. Endogenous pathway (class I MHC) Endogenous antigen Proteasome Evidence Suggests Different Antigen Processing and Presentation Pathways The immune system typically uses different pathways to eliminate intracellular and extracellular antigens. As a general rule, endogenous antigens (those generated within the cell) are processed in the cytosolic or endogenous pathway and presented on the membrane with class I MHC molecules. Exogenous antigens (those taken up from the extracellular environment by endocytosis) are typically processed in the exogenous pathway and presented on the membrane with class II MHC molecules (Figure 8-15). Experiments carried out by L. A. Morrison and T. J. Braciale provided an excellent demonstration that the antigenic peptides presented by class I and class II MHC molecules follow different routes during antigen processing. To do this, they used a set of T-cell clones specific for an influenza virus antigen; some recognized the antigen presented by MHC class I, and others recognized the same antigen presented by MHC class II. Examining the T-cell responses, they derived the following general principles about the two pathways: • Class I presentation requires internal synthesis of virus protein, as shown by the requirement that the target cell be infected by live virus, and by the inhibition of class I presentation observed when protein synthesis was blocked by the inhibitor emetine. • Class II presentation can occur with either live or replication-incompetent virus; protein synthesis inhibitors had no effect, indicating that new protein synthesis is not a necessary condition of class II presentation. • Class II, but not class I, presentation is inhibited by treatment of the cells with an agent that blocks endocytic processing within the cell (e.g., chloroquine). These studies support the distinction between the processing of exogenous and endogenous antigens. They suggest a preferential, but not absolute, association of exogenous antigens with class II MHC molecules and of endogenous antigens with class I molecules. What they do not show is that both the APC of choice and the antigen of choice can influence the outcome, as will be seen later when these rules are subverted during cross-presentation. In the experiments described above, association of viral antigen with class I MHC molecules required replication of the influenza virus and viral protein synthesis within the target cells, but asso- Exogenous pathway (class II MHC) Peptide Rough endoplasmic reticulum (RER) Class I MHC Class II MHC Golgi complex Exogenous antigen Class I MHC Class II MHC FIGURE 8-15 Overview of endogenous and exogenous pathways for processing antigen. In the endogenous pathway (left), antigens are degraded by the proteasome, converting proteins into smaller peptides. In the exogenous pathway (right), extracellular antigens are engulfed into endocytic compartments where they are degraded by acidic pH-dependent endosomal and lysosomal enzymes. The antigenic peptides from proteasome cleavage and those from endocytic compartments associate with class I or class II MHC molecules respectively, and the MHC-peptide complexes are then transported to the cell membrane. It should be noted that the ultimate fate of most peptides in the cell is neither of these pathways but rather to be degraded completely into amino acids. ciation with class II did not. These findings suggested that the peptides presented by class I and class II MHC molecules are trafficked through separate intracellular compartments; class I MHC molecules interact with peptides derived from cytosolic degradation of endogenously synthesized proteins, class II molecules with peptides derived from endocytic degradation of exogenous antigens. The next two sections examine these two pathways in detail. The Major Histocompatibility Complex and Antigen Presentation The Endogenous Pathway of Antigen Processing and Presentation In eukaryotic cells, protein levels are carefully regulated. Every protein is subject to continuous turnover and is degraded at a rate that is generally expressed in terms of its half-life. Some proteins (e.g., transcription factors, cyclins, and key metabolic enzymes) have very short half-lives. Denatured, misfolded, or otherwise abnormal proteins also are degraded rapidly. Defective ribosomal products are polypeptides that are synthesized with imperfections and constitute a large part of the products that are rapidly degraded. The average half-life for cellular proteins is about 2 days, but many are degraded within 10 minutes. The consequence of steady turnover of both normal and defective proteins is a constant deluge of degradation products within a cell. Most will be degraded to their constituent amino acids and recycled, but some persist in the cytosol as peptides. The cell samples these peptides and presents some on the plasma membrane in association with class I MHC molecules, where cells of the immune system can sample these peptides to survey for foreign proteins. The pathway by which these endogenous peptides are generated for presentation with class I MHC molecules utilizes mechanisms similar to those involved in the normal turnover of intracellular proteins, but exactly how particular proteins are selected for degradation and peptide presentation still remains unclear. Peptides Are Generated by Protease Complexes Called Proteasomes Intracellular proteins are degraded into short peptides by a cytosolic proteolytic system present in all cells, called the proteasome (Figure 8-16a). The large (20S) proteasome is composed of 14 subunits arrayed in a barrel-like structure of symmetrical rings. Many proteins are targeted for proteolysis when a small protein called ubiquitin is attached to them. These ubiquitin-protein conjugates enter the proteasome complex, consisting of the 20S base and an attached 19S regulatory component, through a narrow channel at the 19S end. The proteasome complex cleaves peptide bonds in an ATP-dependent process. Degradation of ubiquitin-protein complexes is thought to occur within the central hollow of the proteasome. The immune system also utilizes this general pathway of protein degradation to produce small peptides for presentation by class I MHC molecules. In addition to the standard 20S proteasomes resident in all cells, a distinct proteasome of the same size can be found in pAPCs and the cells of infected tissues. This distinct proteasome, called the immunoproteasome, has some unique components that can be induced by exposure to interferon- or TNF- (Figure 8-16b). LMP2 and LMP7, genes that are located within the class I region (see Figure 8-8) and are responsive to these cytokines, encode replacement catalytic protein subunits that convert | CHAPTER 8 (a) 285 (b) Protein β1 Ubiquitin conjugation β5 β2 Constitutive proteasome Ub 20S proteasome 19S regulator β1i β5i β2i Immunoproteasome FIGURE 8-16 Cytosolic proteolytic system for degradation of intracellular proteins. (a) Endogenous proteins may be targeted for degradation by ubiquitin conjugation. These proteins are degraded by the 26S proteasome complex, which includes the 20S constitutive proteasome and a 19S regulator. (b) In activated APCs, several proteins in the constitutive proteasome (1, 2, and 5) are replaced by proteins encoded by the LMP genes and specific to the immunoproteasome (1i, 2i, and 5i). This immunoproteasome has increased proteolytic efficiency for creating peptides that can assemble with MHC class I molecules. [Adapted from M. Groettrup et al. 2010. Proteasomes in immune cells: More than peptide producers. Nature Reviews. Immunology 10:73–78. doi:10.1038/nri2687] standard proteasomes into immunoproteasomes, increasing the production of peptides that bind efficiently to MHC class I proteins. The immunoproteasome turns over more rapidly than a standard proteasome, possibly because the increased level of protein degradation in its presence may have consequences beyond the targeting of infected cells. It is possible that in some cases autoimmunity results from increased processing of self-proteins in cells with high levels of immunoproteasomes. Peptides Are Transported from the Cytosol to the RER Insight into the cytosolic processing pathway came from studies of cell lines with defects in peptide presentation by class I MHC molecules. One such mutant cell line, called RMA-S, expresses about 5% of the normal levels of class I MHC molecules on its membrane. Although RMA-S cells synthesize normal levels of class I  chain and 2-microglobulin, few class I MHC complexes appear on the membrane. A clue to the mutation in the RMA-S cell line was the discovery by Townsend and his colleagues that “feeding” these cells peptides restored their level of membrane-associated class I MHC molecules to normal. These investigators suggested that peptides might be required to stabilize the interaction between the class I  chain and 2-microglobulin. The ability PA R T I I I 286 | Adaptive Immunity: Antigen Receptors and MHC to restore expression of class I MHC molecules on the membrane by feeding the cells predigested peptides suggested that the RMA-S cell line might have a defect in peptide transport. Subsequent experiments showed that the defect in the RMA-S cell line occurs in the protein that transports peptides from the cytoplasm into the RER, where class I molecules are synthesized. When RMA-S cells were transfected with a functional gene encoding the transporter protein, the cells began to express class I molecules on the membrane. The transporter protein, designated TAP (for transporter associated with antigen processing), is a membrane-spanning heterodimer consisting of two proteins: TAP1 and TAP2 (Figure 8-17a). In addition to their multiple transmembrane TAP1 (a) TAP2 ATP ATP Cytosol RER membrane RER lumen Cytosol (b) Amino acids Protein Peptides Immunoproteasome ATP ADP + Pi TAP RER lumen Peptide ready to be loaded onto class I MHC molecule FIGURE 8-17 TAP (transporter associated with antigen processing). (a) Schematic diagram of TAP, a heterodimer anchored in the membrane of the rough endoplasmic reticulum (RER). The two chains are encoded by TAP1 and TAP2. The cytosolic domain in each TAP subunit contains an ATP-binding site, and peptide transport depends on the hydrolysis of ATP. (b) In the cytosol, association of 1i, 2i, and 5i (colored spheres) with a proteasome changes its catalytic specificity to favor production of peptides that bind to class I MHC molecules. These peptides are translocated by TAP into the RER lumen, where, in a process mediated by several other proteins, they will associate with class I MHC molecules. segments, the TAP1 and TAP2 proteins each have a domain projecting into the lumen of the RER and an ATP-binding domain that projects into the cytosol. Both TAP1 and TAP2 belong to the family of ATP-binding cassette proteins found in the membranes of many cells, including bacteria. These proteins mediate ATP-dependent transport of amino acids, sugars, ions, and peptides. Peptides generated in the cytosol by the proteasome are translocated by TAP into the RER by a process that requires the hydrolysis of ATP (Figure 8-17b). TAP has affinity for peptides containing 8 to 16 amino acids. The optimal peptide length for class I MHC binding is around 9 amino acids, and longer peptides are trimmed by enzymes present in the ER, such as ERAP (endoplasmic reticulum aminopeptidase). In addition, TAP appears to favor peptides with hydrophobic or basic carboxyl-terminal amino acids, the preferred anchor residues for class I MHC molecules. Thus, TAP is optimized to transport peptides that are most likely to interact with class I MHC molecules. The TAP1 and TAP2 genes map within the class II MHC region, adjacent to the LMP2 and LMP7 genes, and different allelic forms of these genes exist within the population. TAP deficiencies can lead to a disease syndrome that has aspects of both immunodeficiency and autoimmunity (see Clinical Focus Box 8-4). Chaperones Aid Peptide Assembly with MHC Class I Molecules Like other proteins destined for the plasma membrane, the  chain and 2-microglobulin components of the class I MHC molecule are synthesized on ribosomes on the RER. Assembly of these components into a stable class I MHC molecular complex that can exit the RER requires the presence of a peptide in the binding groove of the class I molecule. The assembly process involves several steps and includes the participation of molecular chaperones that facilitate the folding of polypeptides. The first molecular chaperone involved in class I MHC assembly is calnexin, a resident membrane protein of the ER. ERp57, a protein with enzymatic activity, and calnexin associate with the free class I  chain and promote its folding (Figure 8-18). When 2-microglobulin binds to the  chain, calnexin is released and the class I molecule associates with the chaperone calreticulin and with tapasin. Tapasin (TAPassociated protein) brings the TAP transporter into proximity with the class I molecule and allows it to acquire an antigenic peptide. The TAP protein promotes peptide capture by the class I molecule before the peptides are exposed to the luminal environment of the RER. Exoproteases in the ER will act on peptides not associated with class I MHC molecules. One ER aminopeptidase, ERAP1, removes the amino-terminal residue from peptides to achieve optimum class I binding size (see Figure 8-18). ERAP1 has little affinity for peptides shorter than eight amino acids in length. As a consequence of productive The Major Histocompatibility Complex and Antigen Presentation | CHAPTER 8 287 BOX 8-4 CLINICAL FOCUS Deficiencies in TAP Can Lead to Bare Lymphocyte Syndrome A relatively rare condition known as bare lymphocyte syndrome (BLS) has been recognized for more than 20 years. The lymphocytes in BLS patients express MHC molecules at below-normal levels and, in some cases, not at all. Type 1 BLS is caused by a deficiency in MHC class I molecules; in type 2 BLS, expression of class II molecules is impaired. The pathogenesis of type 1 BLS underscores the importance of the class I family of MHC molecules in their dual roles of preventing autoimmunity as well as defending against pathogens. One study identified a group of patients with type 1 BLS caused by defects in TAP1 or TAP2 genes. As described in this chapter, TAP proteins are necessary for the loading of peptides onto class I molecules, a step that is essential for expression of class I MHC molecules on the cell surface. Lymphocytes in individuals with TAP deficiency express levels of class I molecules significantly lower than those of normal controls. Other cellular abnormalities include increased numbers of NK and T cells and decreased levels of CD8  T cells. As we will see, the disease manifestations are reasonably well explained by these deviations in the levels of certain cells involved in immune function. In early life, the TAP-deficient individual suffers frequent bacterial infections of the upper respiratory tract and in the second decade begins to experience chronic infection of the lungs. It is thought that a postnasal-drip syndrome common in younger patients promotes the bacterial lung infections in later life. Noteworthy is the absence of susceptibility to severe viral infection, which is common in immunodeficiencies with T-cell involvement (see Chapter 18). Bronchiectasis (dilation of the bronchial tubes) often occurs, and recurring infections can lead to lung damage that may be fatal. The most characteristic mark of the deficiency is the occurrence of necrotizing skin lesions on the extremities and the midface (Figure 1). These lesions ulcerate and may cause disfigurement. The skin lesions are probably due to activated NK cells and T cells; NK cells were isolated from biopsied skin from several patients, supporting this possibility. Normally, the activity of NK cells is limited through the action of killer-cellinhibitory receptors (KIRs), which deliver a negative signal to the NK cell following interaction with class I molecules (see Chapter 13). The lack of class I molecules in BLS patients with TAP deficiency explains the excessive activity of the NK cells. Activation of NK cells further explains the absence of severe viral infections, which are limited by NK and T cells. The best treatment for the characteristic lung infections appears to be a combination of antibiotics and intravenous immunoglobulin. Attempts to limit the skin disease by immunosuppressive regimens, such as steroid treatment or cytotoxic agents, can lead to exacerbation of lesions and are therefore contraindicated. Mutations in the promoter region of TAP that prevent expression of the gene were found for several patients, suggesting the possibility of gene therapy, but the cellular distribution of class I is so widespread that it is not clear what cells would need to be corrected to alleviate all symptoms. FIGURE 1 Necrotizing granulomatous lesions in the midface of a patient with TAPdeficiency syndrome. TAP deficiency leads to a condition with symptoms characteristic of autoimmunity, such as the skin lesions that appear on the extremities and the midface, as well as immunodeficiency that causes chronic sinusitis, leading to recurrent lung infection. [From S. D. Gadola et al., 1999, Lancet 354:1598, and 2000, Clinical and Experimental Immunology 121:173.] PA R T I I I 288 | Adaptive Immunity: Antigen Receptors and MHC RER lumen ERp57 Tapasin ERAP Calreticulin-tapasin– associated class I MHC molecule ERp57 β2-microglobulin Calreticulin Calreticulin Peptide Exit RER + Class I MHC α chain Tapasin Calnexin TAP Peptides Class I MHC molecule Proteasome Cytosol FIGURE 8-18 Assembly and stabilization of class I MHC molecules. Within the rough endoplasmic reticulum (RER) membrane, a newly synthesized class I  chain associates with calnexin, a molecular chaperone, and ERp57 until 2-microglobulin binds to the  chain. The binding of 2-microglobulin releases calnexin and allows binding to calreticulin and to tapasin, which is associated peptide binding, the class I molecule displays increased stability and can dissociate from the complex with calreticulin, tapasin, and ERp57. The class I molecule can then exit from the RER and proceed to the cell surface via the Golgi complex. The Exogenous Pathway of Antigen Processing and Presentation APCs can internalize particulate material by simple phagocytosis (also called “cell eating”), where material is engulfed by pseudopods of the cell membrane, or by receptor-mediated endocytosis, where the material first binds to specific surface receptors. Macrophages and dendritic cells internalize antigen by both processes. Most other APCs, whether professional or not, demonstrate little or no phagocytic activity and therefore typically internalize exogenous antigen only by endocytosis (either receptor-mediated endocytosis or by pinocytosis, “cell drinking”). B cells, for example, internalize antigen very effectively by receptor-mediated endocytosis using their antigenspecific membrane immunoglobulin as the receptor. Peptides Are Generated from Internalized Antigens in Endocytic Vesicles Once an antigen is internalized, it is degraded into peptides within compartments of the endocytic processing pathway. with the peptide transporter TAP. This association promotes binding of an antigenic peptide. Antigens in the ER can be further processed via exopeptidases such as ERAP1, producing fragments ideally suited for binding to class I. Peptide association stabilizes the class I molecule-peptide complex, allowing it to be transported from the RER to the plasma membrane. As the experiment shown in Figure 8-14 demonstrated, internalized antigen takes 1 to 3 hours to traverse the endocytic pathway and appear at the cell surface in the form of class II MHC-peptide complexes. The endocytic antigen processing pathway appears to involve several increasingly acidic compartments, including early endosomes (pH 6.0– 6.5); late endosomes, or endolysosomes (pH 4.5–5.0); and lysosomes (pH 4.5). Internalized antigen progresses through these compartments, encountering hydrolytic enzymes and a lower pH in each compartment (Figure 8-19). Antigen-presenting cells have a unique form of late endosome, the MHC class II-containing compartment (MIIC), in which final protein degradation and peptide loading into MHC class II proteins occurs. Within the compartments of the endocytic pathway, antigen is degraded into oligopeptides of about 13 to 18 residues that meet up with and bind to class II MHC molecules in late endosomes. Because the hydrolytic enzymes are optimally active under acidic conditions (low pH), antigen processing can be inhibited by chemical agents that increase the pH of the compartments (e.g., chloroquine) as well as by protease inhibitors (e.g., leupeptin). The mechanism by which internalized antigen moves from one endocytic compartment to the next has not been conclusively demonstrated. It has been suggested that early endosomes from the periphery move inward to become late endosomes and eventually lysosomes. Alternatively, small The Major Histocompatibility Complex and Antigen Presentation CHAPTER 8 289 the ER and do not transit past the cis-Golgi. However, in cells transfected with both the class II MHC genes and the Ii gene, the class II molecules were localized in the cytoplasmic vesicular structures of the endocytic pathway. The invariant chain contains sorting signals in its cytoplasmic tail that direct the transport of the class II MHC complex from the trans-Golgi network to the endocytic compartments. Antigen Recycling of receptors Clathrincoated vesicle | Early endosome pH 6.0–6.5 Endoplasmic reticulum Golgi complex MHC class II-Ii MIIC Late endosome pH 4.5-5.0 FIGURE 8-19 Generation of antigenic peptides in the exogenous processing pathway. Internalized exogenous antigen moves through several acidic compartments ending in specialized MIIC late endosomes, where it is degraded into peptide fragments which associate with class II MHC molecules transported in vesicles from the Golgi complex. The cell shown here is a B cell, which internalizes antigen by receptor-mediated endocytosis, with the membrane-bound antibody functioning as an antigen-specific receptor. transport vesicles may carry antigens from one compartment to the next. Eventually the endocytic compartments, or portions of them, return to the cell periphery, where they fuse with the plasma membrane. In this way, the surface receptors are recycled. The Invariant Chain Guides Transport of Class II MHC Molecules to Endocytic Vesicles Since APCs express both class I and class II MHC molecules, some mechanism must exist to prevent class II MHC molecules from binding to the antigenic peptides destined for the class I molecules. When class II MHC molecules are synthesized within the RER, these class II  chains associate with a protein called the invariant chain (Ii, CD74). This conserved, non-MHC encoded protein interacts with the class II peptide-binding groove preventing any endogenously derived peptides from binding while the class II molecule is within the RER (Figure 8-20a). The invariant chain also appears to be involved in the folding of the class II  and  chains, their exit from the RER, and the subsequent routing of class II molecules to the endocytic processing pathway from the trans-Golgi network. The role of the invariant chain in the routing of class II molecules has been demonstrated in transfection experiments with cells that lack both the genes encoding class II MHC molecules and the invariant chain. Immunofluorescent labeling of these cells transfected only with class II MHC genes revealed that, in the absence of invariant chain, class II molecules remain primarily in Peptides Assemble with Class II MHC Molecules by Displacing CLIP Recent experiments indicate that most class II MHCinvariant chain complexes are transported from the RER, where they are formed, through the Golgi complex and trans-Golgi network, and then through the endocytic pathway, moving from early endosomes to the MIIC late endosomal compartment in some cases. As the proteolytic activity increases in each successive compartment, the invariant chain is gradually degraded. However, a short fragment of the invariant chain termed CLIP (for class II–associated invariant chain peptide) remains bound to the class II molecule after the majority of the invariant chain has been cleaved within the endosomal compartment. Like antigenic peptide, CLIP physically occupies the peptide-binding groove of the class II MHC molecule, preventing any premature binding of antigen-derived peptide (Figure 8-20b). A nonclassical class II MHC molecule called HLA-DM is required to catalyze the exchange of CLIP with antigenic peptides (see Figure 8-20a). The DM and DM genes are located near the TAP and LMP genes in the MHC complex of humans, with similar genes in mice (see Figure 8-8). Like other class II MHC molecules, HLA-DM is a heterodimer of  and  chains. However, unlike other class II molecules it is relatively nonpolymorphic and is not normally expressed at the cell membrane but is found predominantly within the endosomal compartment. HLA-DM has been found to associate with the class II MHC  chain and to function in removing or “editing” peptides, including CLIP, that associate transiently with the binding groove of classical class II molecules. Peptides that make especially strong molecular interactions with class II MHC, creating long-lived complexes, are harder for HLADM to displace, and thus become the repertoire of peptides that ultimately make it to the cell surface as MHC-peptide complexes. As with class I MHC molecules, peptide binding is required to maintain the structure and stability of class II MHC molecules. Once a peptide has bound, the class II MHC-peptide complex is transported to the plasma membrane, where the neutral pH appears to enable the complex to assume a compact, stable form. Peptide is bound so strongly in this compact form that it is difficult to replace a class II–bound peptide on the membrane with another peptide at physiologic conditions. PA R T I I I 290 (a) | Adaptive Immunity: Antigen Receptors and MHC FIGURE 8-20 Assembly of class II MHC molecules. (a) Within Digested invariant chain Peptides CLIP αβ + MHC Class II Released CLIP – + Invariant chain HLA-DM HLA-DO (b) α1 C N β1 the rough endoplasmic reticulum, a newly synthesized class II MHC molecule binds an invariant chain. The bound invariant chain prevents premature binding of peptides to the class II molecule and helps to direct the complex to endocytic compartments containing peptides derived from exogenous antigens. Digestion of the invariant chain leaves CLIP, a small fragment remaining in the binding groove of the class II MHC molecule. HLA-DM, a nonclassical MHC class II molecule present within the MIIC compartment, mediates exchange of antigenic peptides for CLIP. The nonclassical class II molecule HLA-DO may act as a negative regulator of class II antigen processing by binding to HLA-DM and inhibiting its role in the dissociation of CLIP from class II molecules. (b) Comparison of threedimensional structures showing the binding groove of HLA class II molecules (1, 1) containing different antigenic peptides or CLIP. The red lines show HLA-DR4 complexed with collagen II peptide, yellow lines are HLA-DR1 with influenza hemagglutinin peptide, and dark blue lines are HLA-DR3 associated with CLIP. (N indicates the amino-terminus and C the carboxyl-terminus of the peptides.) No major differences in the structures of the class II molecules or in the conformation of the bound peptides are seen. This comparison shows that CLIP binds the class II molecule in a manner identical to that of antigenic peptides. [Part b from A. Dessen et al., 1997, Immunity 7:473–481; courtesy of Don Wiley, Harvard University.] One additional nonclassical member of the class II MHC family, HLA-DO, like DM, is another relatively nonpolymorphic class II molecule. Unlike other class II molecules, its expression is not induced by IFN- . HLA-DO has been found to act as a negative regulator of antigen binding, modulating the function of HLA-DM and changing the repertoire of peptides that preferentially bind to classical class II molecules. In cells that express both DO and DM, these two molecules strongly associate in the ER and maintain this interaction all the way to the endosomal compartments. Although this interaction has been recognized for many years, the function of this negative regulator and the impact of this changed peptide repertoire is only now being resolved. Originally observed only in B cells and the thymus, the cellular expression profile of HLA-DO has recently been expanded to dendritic cells (DCs), where it is thought to play a role in the maintenance of self-tolerance (discussed further in Chapter 16). This phenomenon was studied by L. K. Denzin and colleagues using diabetes-prone mice engineered to express human HLA-DO. DCs are known to be essential in the establishment of self-tolerance, as well as in the presentation of autoantigens to self-reactive T cells. Selfreactive T cells develop in these non-obese diabetic animals, and these cells ultimately destroy pancreatic  cells, causing diabetes. In this study, the development of diabetes was blocked by the presence of the HLA-DO transgene in DCs. In addition, using specific monoclonal antibodies, they showed that the repertoire of peptides being presented to the autoreactive T cells was significantly altered, resulting in a reduced efficiency in presenting key self-peptides. This and other work suggest that normal DO expression may play an important role in modulating DM behavior to ensure the presentation of a self-peptide repertoire that encourages tolerance to self-antigens. Interestingly, in wildtype mice and humans, DO expression is down-regulated following DC activation by antigen, releasing HLA-DM to carry out its normal function of encouraging the presentation of a diverse array of peptides, many of which will presumably be derived from the foreign proteins that stimulated these APCs. Figure 8-21 illustrates the endogenous pathway (left side) and compares it with the separate exogenous pathway (right), described above. Whether an antigenic peptide associates with class I or with class II molecules is partially dictated by the mode of entry into the cell, either exogenous or endogenous, by the site of processing, and by the cell type. However, in the next section, we will see that these assignments are not absolute and that exogenous antigens may, in some APCs, be presented by class I antigens via a unique pathway. | The Major Histocompatibility Complex and Antigen Presentation Endogenous pathway (class I MHC) CHAPTER 8 291 Exogenous pathway (class II MHC) Endogenous antigen 1 Endogenous antigen is degraded by proteasome. Proteasome 2 Rough endoplasmic reticulum (RER) 1 Peptide is transported to RER via TAP. TAP ERp57 Calreticulin Invariant chain Class II MHC ␣ and ␤ bind invariant chain, blocking binding of endogenous antigen. Tapasin β2- microglobulin Class I MHC Calnexin Class II MHC 2 MHC complex is routed through Golgi to endocytic pathway compartments. Peptide 3 Class I MHC ␣ chain binds calnexin and ERp57, then β2 microglobulin. Calnexin dissociates. Calreticulin and tapasin, bind. MHC captures peptide, chaperones dissociate. 3 Golgi complex CLIP Digested invariant chain 4 Invariant chain is degraded, leaving CLIP fragment. Exogenous antigen is taken up, degraded, routed to endocytic pathway compartments. 4 Class I MHC–peptide is transported from RER to Golgi complex to plasma membrane. Exogenous antigen 5 HLA-DM mediates exchange of CLIP for antigenic peptide. 6 Class I MHC Class II MHC Class II MHC–peptide is transported to plasma membrane. FIGURE 8-21 Separate antigen-presenting pathways are utilized for endogenous (green) and exogenous (red) antigens. The mode of antigen entry into cells and the site of antigen processing determine whether antigenic peptides associate with class I MHC molecules in the rough endoplasmic reticulum or with class II molecules in endocytic compartments. Cross-Presentation of Exogenous Antigens We have seen that after pAPCs internalize extracellular antigen they can process and present these antigens via the exogenous pathway. Because pAPCs also express costimulatory molecules, engagement of CD4 T cells with these class II MHC-peptide complexes can lead to activation of T helper responses. On the other hand, infected cells will generally process and present cytosolic peptides via the endogenous pathway, leading to class I MHC-peptide complexes on their surface. However, unless these infected cells are pAPCs, they will not express the costimulatory molecules necessary to activate naïve CD8 T cells. This leaves us with a dilemma: how does the immune system activate naïve CD8 T cells to eliminate 292 PA R T I I I | Adaptive Immunity: Antigen Receptors and MHC intracellular microbes unless a professional antigenpresenting cell happens to become infected? And if a pAPC is not harboring an intracellular infection, how does this cell present pathogens, such as viruses that have been engulfed from extracellular sources in ways that will activate the needed CTL responses (usually mediated by the endogenous pathway)? The answer to this dilemma is a process called crosspresentation. In some instances, APCs will divert antigen obtained by endocytosis (exogenous antigen) to a pathway that leads to class I MHC loading and peptide presentation to CTLs (like in the endogenous pathway)—in other words, crossing the two pathways. First reported by Michael Bevan and later described in detail by Peter Cresswell and colleagues, the phenomenon of cross-presentation requires that internalized antigens that would normally be handled by the exogenous pathway leading to class II MHC presentation somehow become redirected to a class I peptide loading pathway. When this form of antigen presentation leads to the activation of CTL responses, it is referred to as cross-priming; when it leads to the induction of tolerance in these CD8 T cells it is called cross-tolerance. Dendritic Cells Appear to Be the Primary Cross-Presenting Cell Type The concept of cross-presentation was first recognized as early as 1976, but the relevant cell types and mechanisms of action remained a mystery until more recently. New studies, focused on subsets of DCs that are now recognized to be the most efficient of the cross-presenters, are just beginning to shed light on this process. For example, in vivo depletion or inactivation of DCs compromises cross-presentation. The most potent of the cross-presenting DCs appear to reside in secondary lymphoid organs, where they are believed to receive antigen from extracellular sources by handoff from migrating APCs or from dying infected cells. A few other cells types, such as B cells, macrophages, neutrophils, and mast cells, have been found capable of cross-presentation in vitro, although evidence that they cross-present in vivo or that they have the ability to prime CTL responses (i.e., activate naïve CD8 T cells) is still lacking. Mechanisms and Functions of Cross-Presentation Two possible models have been proposed for how DCs accomplish cross-presentation. The first hypothesizes that cross-presenting cells possess special antigen-processing machinery that allows loading of exogenously derived peptides onto class I MHC molecules. The second theory postulates specialized endocytosis machinery that can send internalized antigen directly to an organelle (such as a phagosome or early endosome), where peptides from those antigens are then loaded onto class I MHC mole- cules using conventional machinery. These two proposed mechanisms are not mutually exclusive, and evidence for both exists. The means by which antigen achieves the crossover from its exogenous or extracellular origins to the endogenous pathway is unresolved. However, in many instances cross-presented antigen has been found to enter the cytoplasm. In the early 1990s, it was shown that bead-conjugated antigens captured by cross-presenting DCs could reach the cytosol of these cells. Later, M. L. Lin and colleagues showed that exogenously added cytochrome c, which causes programmed cell death when present in the cytosol, could induce cross-presenting DCs to undergo apoptosis. It is now proposed that retro-translocation, or movement of endocytosed proteins out of endocytic compartments and into the cytosol, can occur via TAP molecules present in these endocytic membranes—a sort of backflow out of endocytic vesicles. However, TAP-independent cross-presentation of some antigens has also been observed, suggesting multiple mechanisms may exist for targeting externally derived peptides to class I molecules. The ability of some DCs to cross-present antigens has great advantage for the host. It allows these APCs to capture virus from the extracellular environment or from dying cells, process these viral antigens, and activate CTLs that can attack virus-infected cells, inhibiting further spread of the infection. One outstanding question remains: if this pathway is so important, why don’t all APCs cross-present? The answer may lie in the fact that cross-presenting cells could quickly become targets of lysis themselves. Another dilemma concerns how cross-presenting DCs handle self-peptides presented in MHC class I. Shouldn’t these cells break tolerance by presenting extracellularly-derived self-peptides to CD8 T cells, activating anti-self CTL responses? To help explain how cross-presentation is regulated and tolerance is maintained, scientists have proposed that DCs might first need to be “licensed” before they can crosspresent. The cell type postulated to supply this licensing role is activated CD4 T cells. The way this is believed to work is that, first, the classical exogenous pathway of antigen processing in DCs leads to presentation of antigen to CD4 T cells via class II, leading to activation of these cells (Figure 8-22a). These activated helper cells might then return the favor by inducing costimulatory molecules in the DC and by cytokine secretion (e.g., IL-2), supplying a “second opinion” that, respectively, licenses the DC to present internalized antigens via MHC class I and helps activate naïve CD8 T cells (Figure 8-22b). This requirement for licensing by a TH cell could help avoid accidental induction of CTLs to nonpathogenic antigens or self-proteins. If this is the case, any unlicensed DCs that cross-present antigens may serve the opposite and equally important purpose. Since they have not received the go-ahead from the TH cell to activate CTL responses, they may instead induce tolerance in the CD8 T cells they encounter, helping to dampen reactivity to these antigens. The Major Histocompatibility Complex and Antigen Presentation (a) DC licensing by TH cell Dendritic cell Exogenous antigen Exogenous pathway Class II MHC CD4 CD3 TCR TH cell CD40 CD40L (b) DC cross-presentation and activation of CTL Cross-presenting dendritic cell Exogenous antigen TLR Crossover pathway Class I MHC CD8 CD3 CD80/CD86 CD28 IL-2 Naïve TC cell FIGURE 8-22 Activation of naïve Tc cells by exogenous antigen requires DC licensing and cross-presentation. (a) Dendritic cells (DCs) first internalize and process antigen through the exogenous pathway, presenting to CD4 TH cells via MHC class II molecules and activating these cells through, among other things, CD40-CD40L engagement. (b) These activated TH cells can then serve as a bridge to help activate CTL responses; they provide local IL-2 and they in turn license the DC to cross-present internalized antigen in MHC class I, up-regulate costimulatory molecules, and down-regulate their inhibitory counterparts. DC licensing creates an ideal situation for the stimulation of antigen-specific CD8 T cell responses. When the TLRs on these DCs are engaged, this further activates these cells, providing added encouragement for cross-presentation. Dashed arrows indicate antigen directed for cross-presentation. [Adapted from Kurts et al., 2010, Nature Reviews Immunology 10:403.] Presentation of Nonpeptide Antigens To this point, the discussion of the presentation of antigens has been limited to protein antigens and their presentation by classical class I and II MHC molecules. It is well known | CHAPTER 8 293 that some nonprotein antigens are also recognized by T cells, and in the 1980s T-cell proliferation was detected in the presence of nonprotein antigens derived from infectious agents. Many reports now indicate that various types of T cells (expressing as well as  T–cell receptors) can react against lipid antigens, such as mycolic acid, derived from well-known pathogens, such as Mycobacterium tuberculosis. These antigens are presented by members of the CD1 family of nonclassical class I molecules. CD1 molecules share structural similarity to classical MHC class I but overlap functionally with MHC class II. Five human CD1 genes are known; all are encoded on a chromosome separate from the classical class I molecules and display very limited polymorphism. Much like classical MHC class I, CD1 proteins are formed by a transmembrane heavy chain, composed of three extracellular  domains, which associates noncovalently with 2-microglobulin. In terms of trafficking and expression profile, however, most CD1 molecules resemble MHC class II proteins, moving intracellularly to endosomal compartments, where they associate with exogenous antigen. Like MHC class II molecules, CD1 proteins are expressed by many immune cell types, including thymocytes, B cells, and DCs, although some members of the family have also been found on hepatocytes and epithelial cells. Many different lipid or lipid-linked structures have been found to associate non-covalently with CD1 molecules. In general, most of these are glycolipid or lipoprotein antigens, where the lipid moiety fits into deep pockets within the CD1 binding groove and the hydrophilic head group remains exposed, allowing recognition by T cells. Crystal structures have demonstrated that CD1 contains a binding groove that is both deeper and narrower than classical MHC molecules. These grooves are lined with nonpolar amino acids, which can easily accommodate lipid structures. Some of these pockets correspond roughly with the antigen-binding pockets of classical MHC class I and allow antigen access to a deep groove through a narrow opening—more like a foot sliding into a shoe (Figure 8-23a). Figure 8-23b illustrates a comparison of the binding groove of CD1b holding a lipid antigen with a classical class I molecule complexed with peptide antigen. The relatively nonspecific manner of antigen association with the CD1 binding groove, which relies primarily on many hydrophobic interactions, probably accounts for the diversity of self- and foreign antigens that can be presented by these nonclassical class I molecules. A variety of T cells are known to bind these nonclassical MHC molecules. It is now hypothesized that short-chain self-lipids with relatively low affinity are loaded onto CD1 molecules in the ER, shortly after translation, and allow proper CD1 protein folding. These self-antigen loaded CD1 molecules then travel to the cell surface, where in some cases exogenous lipids may be exchanged with these low-affinity self-antigens. Like with MHC class II molecules, after endocytosis and movement to lower pH environments, CD1 proteins are believed to 294 PA R T I I I | Adaptive Immunity: Antigen Receptors and MHC (a) A C F T β2-microglobulin (b) FIGURE 8-23 Lipid antigen binding to the CD1 molecule. (a) Cartoon of the deep binding pockets (A, C, and F) and the deep binding tunnel (T) that accommodate lipid antigens in a foot-in-shoe fashion in the members of the CD1 family of nonclassical class I molecules. (b) Ribbon diagram of the binding groove or pocket of human CD1b complexed with lipid (left) compared with a classical class I molecule binding peptide antigen (right). [Part b PDB IDs 1GZQ (left) and 1HHG (right).] exchange their self peptide, low-affinity binding partners for exogenously derived long-chain, high-affinity lipid antigens resident in phagolysosomes or late endosomes. These newly loaded CD1 molecules then return to the cell surface for recognition by CD1-restricted T cells. Recently constructed transgenic mice that express human CD1 molecules will likely lead the way to new murine models aimed at better defining the role of CD1 restricted T cells in protection from infection. The Major Histocompatibility Complex and Antigen Presentation | CHAPTER 8 295 S U M M A R Y ■ ■ ■ ■ ■ ■ ■ The major histocompatibility complex (MHC) encodes class I and II molecules, which function in antigen presentation to T cells, and class III molecules, which have diverse functions. Class I MHC molecules consist of a large glycoprotein  chain encoded by an MHC gene, and 2-microglobulin, a protein with a single domain that is encoded elsewhere. Class II MHC molecules are composed of two noncovalently associated glycoproteins, the  and  chains, encoded by separate MHC genes. MHC genes are tightly linked and generally inherited as a unit from parents; these linked units are called haplotypes. MHC genes are polymorphic (many alleles exist for each gene in the population), polygenic (several different MHC genes exist in an individual), and codominantly expressed (both maternal and paternal copies). MHC alleles influence the fragments of antigen that are presented to the immune system, thereby influencing susceptibility to a number of diseases. Class I molecules are expressed on most nucleated cells; class II molecules are restricted to B cells, macrophages, and dendritic cells (pAPCs). ■ ■ ■ ■ ■ ■ In most cases, class I molecules present processed endogenous antigen to CD8 TC cells and class II molecules present processed exogenous antigen to CD4 TH cells. Endogenous antigens are degraded into peptides within the cytosol by proteasomes, assemble with class I molecules in the RER, and are presented on the membrane to CD8 TC cells. This is the endogenous processing and presentation pathway. Exogenous antigens are internalized and degraded within the acidic endocytic compartments and subsequently combine with class II molecules for presentation to CD4 TH cells. This is the exogenous processing and presentation pathway. Peptide binding to class II molecules involves replacing a fragment of invariant chain in the binding groove by a process catalyzed by nonclassical MHC molecule HLA-DM. In some cases, exogenous antigens in certain cell types (mainly DCs) can gain access to class I presentation pathways in a process called cross-presentation. Presentation of nonpeptide (lipid and lipid-linked) antigens derived from pathogens involves the nonclassical class I–like CD1 molecules. R E F E R E N C E S Amigorena, S., and A. Savina. 2010. Intracellular mechanisms of antigen cross presentation in dendritic cells. Current Opinion in Immunology 22:109. Margulies, D. 1999. The major histocompatibility complex. In Fundamental Immunology, 4th ed. W. E. Paul, ed. LippincottRaven, Philadelphia. Brown, J. H., et al. 1993. Three-dimensional structure of the human class II histocompatibility antigen HLA-DR1. Nature 364:33. Moody, D. B., D. M. Zajonc, and I. A. Wilson. 2005. Anatomy of CD1-lipid antigen complexes. Nature Reviews in Immunology 5:387. Doherty, P. C., and R. M. Zinkernagel. 1975. H-2 compatibility is required for T-cell mediated lysis of target cells infected with lymphocytic choriomeningitis virus. Journal of Experimental Medicine 141:502. Fahrer, A. M., et al. 2001. A genomic view of immunology. Nature 409:836. Gadola, S. D., et al. 2000. TAP deficiency syndrome. Clinical and Experimental Immunology 121:173. Horton, R., et al. 2004. Gene map of the extended human MHC. Nature Reviews Genetics 5:1038. International Human Genome Sequencing Consortium. 2001. Initial sequencing and analysis of the human genome. Nature 409:860. Rock, K. L., et al. 2004. Post-proteosomal antigen processing for major histocompatibility complex class I presentation. Nature Immunology 5:670. Rothenberg, B. E., and J. R. Voland. 1996. Beta 2 knockout mice develop parenchymal iron overload: A putative role for class I genes of the major histocompatibility complex in iron metabolism. Proceedings of the National Academy of Science U.S.A. 93:1529. Rouas-Freiss, N., et al. 1997. Direct evidence to support the role of HLA-G in protecting the fetus from maternal uterine natural killer cytolysis. Proceedings of the National Academy of Science USA. 94:11520. Strominger, J. L. 2010. An alternative path for antigen presentation: Group 1 CD1 proteins. Journal of Immunology 184(7):3303 Kelley, J., et al. 2005. Comparative genomics of major histocompatibility complexes. Immunogenetics 56:683. Wearsch, P. A., and P. Cresswell. 2008. The quality control of MHC class I peptide loading. Current Opinion in Cell Biology 20(6):624. Kurts, C., B. W. S. Robinson, and P. A. Knolle. 2010. Cross-priming in health and disease. Nature Reviews in Immunology 10:403. Yaneva, R., et al. 2010. Peptide binding to MHC class I and II proteins: New avenues from new methods. Molecular Immunology 47:649. Li, X. C., and M. Raghavan. 2010. Structure and function of major histocompatibility complex class I antigens. Current Opinions in Organ Transplantation 15:499. Madden, D. R. 1995. The three-dimensional structure of peptide-MHC complexes. Annual Review of Immunology 13:587. Yi, W., et al. 2010. Targeted regulation of self-peptide presentation prevents type I diabetes in mice without disrupting general immunocompetence. Journal of Clinical Investigation 120(4):1324. PA R T I I I 296 | Adaptive Immunity: Antigen Receptors and MHC map of the human MHC from the Horton et al. reference cited above. Useful Web Sites www.bioscience.org This Frontiers in Bioscience site includes in the Research Tools section a database that lists gene knockouts. Information is available on studies of the consequences of targeted disruption of MHC molecules and other component molecules, including 2-microglobulin and the class II invariant chain. www.bshi.org.uk The British Society for Histocompatibility and Immunogenetics home page contains information on tissue typing and transplantation and links to worldwide sites related to the MHC. www.ebi.ac.uk/imgt/hla The International ImMunoGeneTics (IMGT) database section contains links to sites with information about HLA gene structure and genetics and up-to-date listings and sequences for all HLA alleles officially recognized by the World Health Organization HLA nomenclature committee. www.hla.alleles.org A Web site maintained by the HLA Informatics Group based at the Anthony Nolan Trust in the United Kingdom, with up-to-date information on the numbers of HLA alleles and proteins. www.nature.com/nrg/journal/v5/n12/poster/ MHCmap/index.html A poster with an extended gene S T U D Y Q U E S T I O N S CLINICAL FOCUS QUESTION Patients with TAP deficiency have partial immunodeficiency as well as autoimmune manifestations. How do the profiles for patients’ immune cells explain the partial immunodeficiency? Why is it difficult to design a gene therapy treatment for this disease, despite the fact that a single gene defect is implicated? MHC molecules expressed on the membrane of the transfected L cells Transfected gene Dk Db Kk Kb IAk IAb None 1. Indicate whether each of the following statements is true or false. If you think a statement is false, explain why. a. A monoclonal antibody specific for 2-microglobulin b. c. d. e. f. g. can be used to detect both class I MHC K and D molecules on the surface of cells. Antigen-presenting cells express both class I and class II MHC molecules on their membranes. Class III MHC genes encode membrane-bound proteins. In outbred populations, an individual is more likely to be histocompatible with one of its parents than with its siblings. Class II MHC molecules typically bind to longer peptides than do class I molecules. All nucleated cells express class I MHC molecules. The majority of the peptides displayed by class I and class II MHC molecules on cells are derived from selfproteins. d k 2. You cross a BALB/c (H-2 ) mouse with a CBA (H-2 ) mouse. What MHC molecules will the F1 progeny express on (a) its liver cells and (b) its macrophages? 3. To carry out studies on the structure and function of the class I MHC molecule Kb and the class II MHC molecule IAb, you decide to transfect the genes encoding these proteins into a mouse fibroblast cell line (L cell) derived from the C3H strain (H-2k). L cells do not normally function as antigen-presenting cells. In the following table, indicate which of the listed MHC molecules will () or will not () be expressed on the membrane of the transfected L cells. Kb IA␣b IA␤b IA␣b and IA␤b s 4. The SJL mouse strain, which has the H-2 haplotype, has a deletion of the IE locus. a. List the classical MHC molecules that are expressed on the membrane of macrophages from SJL mice. k b. If the class II IE and IE genes from an H-2 strain are transfected into SJL macrophages, what additional classical MHC molecules would be expressed on the transfected macrophages? 5. Draw diagrams illustrating the general structure, including the domains, of class I MHC molecules, class II MHC molecules, and membrane-bound antibody on B cells. Label each chain and the domains within it, the antigen-binding regions, and regions that have the immunoglobulin-fold structure. 6. One of the characteristic features of the MHC is the large number of different alleles at each locus. a. Where are most of the polymorphic amino acid resi- dues located in MHC molecules? What is the significance of this location? b. How is MHC polymorphism thought to be generated? 7. As a student in an immunology laboratory class, you have been given spleen cells from a mouse immunized with the LCM virus (LCMV). You determine the antigen-specific The Major Histocompatibility Complex and Antigen Presentation functional activity of these cells with two different assays. In assay 1, the spleen cells are incubated with macrophages that have been briefly exposed to LCMV; the production of interleukin 2 (IL-2) is a positive response. In assay 2, the spleen cells are incubated with LCMV-infected target cells; lysis of the target cells represents a positive response in this assay. The results of the assays using macrophages and target cells of different haplotypes are presented in the table below. Note that the experiment has been set up in a way to exclude alloreactive responses (reactions against nonself MHC molecules). a. The activity of which cell population is detected in each b. c. d. e. of the two assays? The functional activity of which MHC molecules is detected in each of the two assays? From the results of this experiment, which MHC molecules are required, in addition to the LCM virus, for specific reactivity of the spleen cells in each of the two assays? What additional experiments could you perform to unambiguously confirm the MHC molecules required for antigen-specific reactivity of the spleen cells? Which of the mouse strains listed in the table could have been the source of the immunized spleen cells tested in the functional assays? Give your reasons. MHC haplotype of macrophages and virus-infected target cells Response of spleen cells CHAPTER 8 297 10. The hypothetical allelic combination HLA-A99 and HLA- B276 carries a relative risk of 200 for a rare, and yet unnamed, disease that is fatal to preadolescent children. a. Will every individual with A99/B276 contract the disease? b. Will everyone with the disease have the A99/B276 combination? c. How frequently will the A99/B276 allelic combination be observed in the general population? Do you think that this combination will be more or less common than predicted by the frequency of the two individual alleles? Why? 11. Explain the difference between the terms antigen-presenting cell and target cell, as they are commonly used in immunology. 12. Define the following terms: a. b. c. d. e. f. Self-MHC restriction Antigen processing Endogenous antigen Exogenous antigen Anchor residues Immunoproteasome 13. Ignoring cross-presentation, indicate whether each of the cell components or processes listed is involved in the processing and presentation of exogenous antigens (EX), endogenous antigens (EN), or both (B). Briefly explain the function of each item. a. b. c. d. e. f. ______ Class I MHC molecules ______ Class II MHC molecules ______ Invariant (Ii) chains ______ Lysosomal hydrolases ______ TAP1 and TAP2 proteins ______ Transport of vesicles from the RER to the Golgi complex ______ Proteasomes ______ Phagocytosis or endocytosis ______ Calnexin ______ CLIP ______ Tapasin Mouse strain used as source of macrophages and target cells K IA IE D C3H k k k k   BALB/C d d d d   d/k d/k d/k d/k   A.TL s k k d   a. If chloroquine is added to the incubation mixture, pre- B10.A (3R) b b b d   B10.A (4R) k k — b   sentation of the native protein is inhibited, but the peptide continues to induce TH-cell activation. Explain why this occurs. b. If chloroquine addition is delayed for 3 hours, presentation of the native protein is not inhibited. Explain why this occurs. (BALB/C X B10.A)F1 IL-2 production in response Lysis of to LCMVLCMVpulsed infected macrophages cells (assay 1) (assay 2) | 8. A TC-cell clone recognizes a particular measles virus pep- tide when it is presented by H-2Db. Another MHC molecule has a peptide-binding groove identical to the one in H-2Db but differs from H-2Db at several other amino acids in the 1 domain. Predict whether the second MHC molecule could present this measles virus peptide to the TC-cell clone. Briefly explain your answer. 9. Human red blood cells are not nucleated and do not express any MHC molecules. Why is this property fortuitous for blood transfusions? g. h. i. j. k. 14. Antigen-presenting cells have been shown to present lyso- zyme peptide 46–61 together with the class II IAk molecule. When CD4 TH cells are incubated with APCs and native lysozyme or the synthetic lysozyme peptide 46–61, TH-cell activation occurs. 15. Cells that can present antigen to TH cells have been classified into two groups: professional and nonprofessional APCs. a. Name the three types of professional APCs. For each type, indicate whether it expresses class II MHC molecules and a costimulatory signal constitutively or must be activated before doing so. b. Give three examples of nonprofessional APCs. When are these cells most likely to function in antigen presentation? PA R T I I I 298 | Adaptive Immunity: Antigen Receptors and MHC 16. Predict whether TH-cell proliferation or CTL-mediated 21. Define the following terms and give examples using the cytolysis of target cells will occur with the following mixtures of cells. The CD4 TH cells are from lysozymeprimed mice, and the CD8 CTLs are from influenza-infected mice. Use R to indicate a response and NR to indicate no response. human HLA locus: polygeny, polymorphism, and codominant expression. How exactly does each contribute to ensuring that a diversity of antigens can be presented by each individual? k b. c. d. e. f. phages ______ H-2k TH cells  lysozyme-pulsed H-2b/k macrophages ______ H-2k TH cells  lysozyme-primed H-2d macrophages ______ H-2k CTLs  influenza-infected H-2k macrophages ______ H-2k CTLs  influenza-infected H-2d macrophages ______ H-2d CTLs  influenza-infected H-2d/k macrophages 17. Molecules of the CD1 family were recently shown to pres- ent nonpeptide antigens. a. What is a major source of nonpeptide antigens? b. Why are CD1 molecules not classified as members of the MHC family even though they associate with 2-microglobulin? c. What evidence suggests that the CD1 pathway is different from that utilized by classical class I MHC molecules? 18. A slide of macrophages was stained by immunofluores- cence using a monoclonal antibody for the TAP1/TAP2 complex. Which of the following intracellular compartments would exhibit positive staining with this antibody? a. b. c. d. Cell surface Endoplasmic reticulum Golgi apparatus The cells would not be positive because the TAP1/TAP2 complex is not expressed in macrophages. 19. HLA determinants are used not only for tissue typing of transplant organs, but also as one set of markers in paternity testing. Given the following phenotypes, which of the potential fathers is most likely the actual biological father? Indicate why each could or could not be the biological father. 22. What is the cellular phenotype that results from inactivation or mutation in both copies of the gene for the invariant chain? 23. In general terms, describe the process of cross-presenta- tion. Which cell types are most likely to be involved in cross-presentation, and what unique role does this process play in the activation of naïve CD8 T cells? ANALYZE THE DATA Smith and colleagues (2002, Journal of Immunology 169:3105) examined the ability of two peptides to bind two different allotypes—Ld and Lq—of the mouse class I MHC molecule. They looked at (a) the binding of a murine cytomegalovirus peptide (MCMV; amino acid sequence YPHFMPTNL) and (b) a synthesized peptide, tum P91A 14–22 (TQNHRALDL). Before and after pulsing the cells with the target peptides, the investigators measured the amount of peptide-free MHC molecules on the surface of cells expressing either Ld, Lq, or a mutant Ld in which tryptophan had been mutated to arginine at amino acid 97 (Ld W97R). The ability of the peptide to decrease the relative number of open forms on the cell surface reflects increased peptide binding to the class I MHC molecules. Answer the following questions based on the data and what you have learned from reading this book. (a) 80 Relative amount of open forms on cell surface k a. ______ H-2 TH cells  lysozyme-pulsed H-2 macro- (b) 60 70 50 No exogenous peptide + MCMV peptide 60 50 No exogenous peptide + tum_ P91A 14–22 peptide 40 40 30 0.77 30 20 20 10 0 10 0.04 Ld 0.08 Lq 0.11 LdW97R 0.71 0.02 Ld Lq LdW97R a. Are there allotypic differences in the binding of peptide HLA-A HLA-B HLA-C Offspring A3, 43 B54, 59 C5, 8 Biological Mother A3, 11 B59, 78 C8, 8 Potential Father 1 A3, 33 B54, 27 C5, 5 Potential Father 2 A11, 43 B54, 27 C5, 6 Potential Father 3 A11, 33 B59, 26 C6, 8 20. Assuming the greatest degree of heterozygosity, what is the maximum number of different classical HLA class I molecules that one individual can employ to present antigens to CD8 T cells? Assuming no extra beta chain genes, what is the maximum number of different class II molecules (isoforms) that one individual can use to present antigen to CD4 T cells? by native class I MHC molecules on the cell surface? b. Is there a difference in the binding of MCMV peptide to Ld after a tryptophan (W) to arginine (R) mutation in the Ld molecule at position 97? Explain your answer. 2 c. Is there a difference in the binding of tum P91A 14–22 d peptide to L after a tryptophan (W) to arginine (R) mutation in the Ld molecule at position 97? Explain your answer.  d. If you wanted to successfully induce a CD8 T-cell response 2 against tum P91A 14–22 peptide, would you inject a mouse that expresses Lq or Ld? Explain your answer. e. The T cells generated against the MCMV peptide have a different specificity than the T cells generated against the tum2 P91A 14–22 peptide when Ld is the restricting MHC molecule. Explain how different peptides can bind the same Ld molecule yet restrict/present peptide to T cells with different antigen specificities. 9 T-Cell Development T cells initiate the adaptive immune response by interacting, via their T-cell receptors, with MHC/peptide complexes on antigen-presenting cells that have been exposed to pathogens. The range of pathogens to which we are exposed is considerable, and, as described in Chapter 7, vertebrates have evolved a remarkable mechanism to generate a comparable range of T-cell (and B-cell) receptor specificities. Each of the several million T cells circulating in the body expresses a distinct T-cell receptor. The generation of this diverse population, with its diverse receptor repertoire, takes place in the thymus, an organ both required for and dedicated to the development of T cells. Immature cells entering the thymus from the bone marrow express no mature lymphocyte features and no antigen receptors; those leaving the thymus are mature T cells that are diverse in their receptor specificities, and are both tolerant to self and restricted to self-MHC. How do we generate such a diverse self-restricted and self-tolerant group of T cells? This question has intrigued immunologists for decades and has inspired and required much experimental ingenuity. The innovation has paid off and, although our understanding is not comprehensive, we now have a fundamental appreciation of the remarkable strategies taken by the thymus, the nursery of immature T cells or thymocytes, to produce a functional, safe, and useful T-cell repertoire. In this chapter we describe these strategies and divide T-cell development into two clusters of events: early thymocyte development, during which a dizzyingly diverse TCR⫹ population of immature T cells is generated, and selection events that depend on TCR interactions to shape this population so that only those cells that are self-restricted and selftolerant will leave to populate the periphery (Overview Figure 9-1). Early thymocyte development is T-cell receptor independent and brings cells through uncommitted CD4⫺CD8⫺ (double negative, DN) stages to the T-cell receptor-expressing, Thymocytes in the cortex of the thymus. ■ Early Thymocyte Development ■ Positive and Negative Selection ■ Lineage Commitment ■ Exit from the Thymus and Final Maturation ■ Other Mechanisms That Maintain Self-Tolerance ■ Apoptosis CD4⫹CD8⫹ (double positive, DP) stage. The specific events in this early stage include: 1. commitment of hematopoietic precursors to the T cell lineage, 2. the initiation of antigen receptor gene rearrangements, and 3. the selection and expansion of cells that have successfully rearranged one of their T-cell receptor genes (-selection). The second phase of T-cell development is largely dependent on T-cell receptor interactions and brings cells to maturity from the CD4⫹CD8⫹ stage to the CD4⫹ or CD8⫹ single positive (SP) stage. The events in this last phase of development include: 1. positive selection, selection for those cells whose T-cell receptors respond to self-MHC, 2. negative selection, selection against those cells whose T-cell receptors react strongly to self-peptide/MHC combinations, and 3. lineage commitment, commitment of thymocytes to effector cell lineages, including CD4⫹ helper or CD8⫹ cytotoxic populations. 299 9-1 OVERVIEW FIGURE Development of T cells in the mouse Hematopoietic stem cell (HSC) Marrow Hematopoietic precursor Blood migration T-cell precursor Dendritic cell, B cell, NK DN1 TCR␤ locus rearrangement, proliferation DN2 Thymic cortex Early thymocyte development CD4⫺CD8⫺ (DN) T-cell commitment DN3 ␤-selection, proliferation DN4 TCR ␣ locus rearrangement Double positive CD4⫹CD8⫹ (DP) Death by neglect Positive and negative selection CD8⫹ CD4⫹ Single positive (SP) Thymic medulla Blood Negative selection CD8⫹ ␥␦ T cell; few further changes in surface phenotype Other lineages, including TREGs, TH17, IEL, NK1.1 CD4⫹ migration ␥␦ T cell Peripheral tissues CD8⫹ T-cell precursors from the bone marrow travel to the thymus via the bloodstream, undergo development to mature T cells, and are exported to the periphery, where they can undergo antigeninduced activation and differentiation into effector cells and memory cells. Each stage of development occurs in a specific microenvironment and is characterized by specific intracellular events and distinctive cell-surface markers. The most immature thymocytes are CD4⫺CD8⫺ (double negative, DN) and pass through several stages (DN1-DN4) during which they commit to the T-cell lineage and begin to rearrange their T-cell receptor (TCR) gene loci. Those that successfully rearrange their TCR chain proliferate, initiate rearrangement of their TCR␣ chains, and become CD4⫹CD8⫹ (double positive, DP) thymocytes, which dominate the 300 CD4⫹ thymus. DP thymocytes undergo negative and positive selection in the thymic cortex (see text and Overview Figure 9-5 for details). Positively selected thymocytes continue to mature and migrate to the medulla, where they are subject to another round of negative selection to self-antigens that include tissue-specific proteins. Mature T cells express either CD4 or CD8 (single positive, SP) and leave the thymus with the potential to initiate an immune response. Although most thymocytes develop into conventional TCR␣␤ CD4⫹ or CD8⫹ T cells, some DN and DP thymocyte cells develop into other cell lineages, including lymphoid dendritic cells, TCR␥␦ T cells, natural killer T cells (NKT), regulatory T cells (TREG), and intraepithelial lymphocytes (IELs), each of which has a distinct function (see text). T-Cell Development Understanding T-cell development has been a challenge for students and immunologists alike, but it helps to keep in mind the central purpose of the process: to generate a large population of T cells that express a diverse array of receptor specificities that can interact with self-MHC, but do not respond to self proteins. | CHAPTER 9 301 Bone marrow stem cells Early Thymocyte Development T-cell development occurs in the thymus and begins with the arrival of small numbers of lymphoid precursors migrating from the bone marrow and blood into the thymus, where they proliferate, differentiate, and undergo selection processes that result in the development of mature T cells. The precise identity of the bone marrow cell precursor that gives rise to T cells is still debated. However, it is clear that this precursor, which is directed to the thymus via chemokine receptors, retains the potential to give rise to more than one type of cell, including natural killer (NK) cells, dendritic cells (DC), B cells, and even myeloid cells. This precursor only becomes fully committed to the T-cell lineage in the late DN2 stage of T-cell development (see Overview Figure 9-1). Studies revealed that commitment to the T-cell lineage was dependent on a receptor, Notch, which had been classically associated with embryonic cell development. Notch, in fact, regulates the decision of a lymphoid precursor to become a T versus a B lymphocyte. When a constitutively active version of Notch1, one of four versions of Notch, is overexpressed in hematopoietic cells, T cells rather than B cells develop in the bone marrow. Reciprocally, when the Notch1 gene is knocked out among hematopoietic precursors, B cells rather than T cells develop in the thymus (!). The importance of Notch in T-cell commitment was underscored by an in vitro system for studying T-cell development. Until recently, we thought that development of T cells from their hematopoietic precursors required the intact architecture of the thymus. Whereas B-cell development could be achieved in vitro using a preparation of bone marrow stem cells grown on stromal cells in the presence of appropriate cytokines, in vitro studies of T-cell development once could only be performed using intact fetal thymic organ culture systems. In 2002, J. C. Zuniga-Pfluker and colleagues demonstrated that T cells could be induced to develop when bone marrow stem cells were cultured on a stromal cell line that expressed a ligand for Notch. As shown in Figure 9-2, growth of hematopoietic stem cells (HSCs) on stromal cells that express Notch ligand drives the development of these multipotent stem cells to the T-cell lineage. This assay system has been invaluable in defining interactions that control early T-cell development and have helped investigators reveal, for instance, that the transcription factor GATA-3 is a critical participant in Notch-mediated T-cell commitment. Stromal cells Notch1⫹ Notch1⫺ Growth factors IL-7, FLT3L 1-2 weeks Harvest T-lineage cells Harvest B-lineage cells FIGURE 9-2 Development of T cells from hematopoietic stem cells on bone marrow stromal cells expressing the Notch ligand. Investigators can induce lymphoid development from hematopoietic stem cells in vitro using a combination of stromal cell lines and soluble cytokines and growth factors, as indicated. Investigators discovered that Notch signaling was the key to inducing development to the T- rather than B-lymphocyte lineage. After transfecting the stromal cell line with a gene encoding the Notch ligand, lymphoid precursors would adopt the T-cell lineage. Otherwise, they would develop into B cells. [Adapted from J. C. Zuniga-Pflucker, 2002, Nature Reviews Immunology 4:67–72.] Thymocytes Progress through Four DoubleNegative Stages T-cell development is elegantly organized, spatially and temporally. Different stages of development take place in distinct microenvironments that provide membrane-bound and soluble signals that regulate maturation. After arriving in the thymus from the bone marrow via blood vessels at the cortico-medullary boundary, T-cell precursors encounter Notch ligands, which are abundantly expressed by the thymic epithelium. Recall that T-cell precursors first travel to the outer cortex where they slowly proliferate, then they pass through the thymic medulla before exiting at the corticomedullary junction (see Figure 2-6). During the time it takes cells to develop in the thymus (1 to 3 weeks), thymocytes pass through a series of stages defined by changes in their cell surface phenotype (see Overview Figure 9-1). The earliest T cells lack detectable CD4 and CD8 and are therefore referred to as double-negative (DN) cells. DN T cells can be subdivided 302 TABLE 9-1 PA R T I V | Adaptive Immunity: Development Double-negative thymocyte development Genotype DN1 c-kit (CD117)⫹⫹, CD44⫹, CD25⫺ ⫹⫹ ⫹ ⫹ Location Description Bone marrow to thymus Migration to thymus TCR␥, ␦, and ␤ chain rearrangement; T-cell lineage commitment DN2 c-kit (CD117) , CD44 , CD25 Subcapsular cortex DN3 c-kit (CD117)⫹, CD44⫺, CD25⫹ Subcapsular cortex Expression of pre-TCR; ␤ selection DN4 c-kit (CD117)low/⫺, CD44⫺, CD25⫺ Subcapsular cortex to cortex Proliferation, allelic exclusion of ␤-chain locus; ␣-chain locus rearrangement begins; becomes DP thymocyte into four subsets (DN1-4) based on the presence or absence of other cell surface molecules, including c-kit (CD117), the receptor for stem cell growth factor; CD44, an adhesion molecule; and CD25, the ␣ chain of the IL-2 receptor (Table 9-1). DN1 thymocytes are the first to enter the thymus and are still capable of giving rise to multiple cell types. They express only c-kit and CD44 (c-kit⫹⫹CD44⫹CD25⫺), but once they encounter the thymic environment and become resident in the cortex, they proliferate and express CD25, becoming DN2 thymocytes (c-kit⫹⫹CD44 ⫹CD25 ⫹). During this critical stage of development, the genes for the TCR ␥, ␦, and ␤ chains begin to rearrange; however, the TCR ␣ locus does not rearrange, presumably because the region of DNA encoding TCR␣ genes is not yet accessible to the recombinase machinery. At the late DN2 stage, T-cell precursors fully commit to the T-cell lineage and reduce expression of both c-kit and CD44. Cells in transition from the DN2 to DN3 (c-kit⫹CD44⫺CD25⫹) stages continue rearrangement of the TCR␥, TCR␦, and TCR␤ chains and make the first major decision in T-cell development: whether to join the TCR␥␦ or TCR␣␤ T-cell lineage. Those DN3 T cells that successfully rearrange their ␤ chain and therefore commit to the TCR␣␤ T-cell lineage lose expression of CD25, halt proliferation, and enter the final phase of their DN stage of development, DN4 (c-kit low/⫺CD44⫺CD25⫺), which mature directly into CD4⫹CD8⫹ DP thymocytes. Thymocytes Can Express Either TCR␣␤ or TCR␥␦ Receptors Vertebrates generate two broad categories of T cells: those that express TCR␣ and ␤ receptor chains and those that express TCR␥ and ␦ receptor chains. TCR␣␤ cells are the dominant participants in the adaptive immune response in secondary lymphoid organs; however, TCR␥␦ cells also play an important role, particularly in protecting our mucosal tissues from outside infection. Both types of cells are generated in the thymus, but how does a cell make the decision to become one or the other? To a large extent, the choice to become a ␥␦ or ␣␤ T cell is dictated by when and how fast the genes that code for each of the four receptor chains successfully rearrange. Recall from Chapter 7 that TCR genes are generated by the shuffling (rearrangement) of V and J (and sometimes D) segments, an event responsible for the vast diversity of receptor specificities. Rearrangement of the ␤, ␥, and ␦ loci begins during the DN2 stage. To become an ␣␤ T cell, a cell must generate a TCR␤ chain—an event that depends on a single in-frame VDJ rearrangement event. To become a ␥␦ cell, however, a thymocyte must generate two functional proteins that depend on two separate in-frame rearrangement events. Probability favors the former fate and, in fact, T cells are at least three times as likely to become TCR␣␤ cells than TCR␥␦ cells. TCR␥␦ T-cell generation is also regulated developmentally. They are the first T cells that arise during fetal development, and provide a very important protective function perhaps even prior to birth. Studies show, for instance, that ␥␦ T cells are required to protect very young mice against the protozoal pathogen that causes coccidiosis. However, production of ␥␦ T cells declines after birth, and the TCR␥␦ T-cell population represents only 0.5% of all mature thymocytes in the periphery of an adult animal (Figure 9-3). Most TCR␥␦ T cells are quite distinct in phenotype and function from conventional TCR␣␤ T cells. Most do not go through the DP stage of thymocyte development and leave the thymus as mature DN T cells. Many emerge from the thymus with the ability to secrete cytokines, a capacity gained by most TCR␣␤ cells only after they encounter antigen in secondary lymphoid tissues. The TCR␥␦ T-cell population also expresses receptors that are not as diverse as TCR␣␤ T cells, and many appear to recognize unconventional antigens, including lipids associated with unconventional MHC molecules. Many take up long-term residence in mucosal tissues and skin and join innate immune cells in providing a first line of attack against invading microbes, as well as the response to cellular stress. T-Cell Development 100 γδ Thymocytes CD3+ cells, % 75 αβ Thymocytes 2. 3. 25 4. 0 14 15 16 17 18 19 Birth Days of gestation Adult FIGURE 9-3 Time course of appearance of ␥␦ thymocytes and ␣␤ thymocytes during mouse fetal development. The graph shows the percentage of cells in the thymus that are double negative (CD4⫺CD8⫺) and bear the ␥␦ T-cell receptor (black) or are double positive (CD4⫹CD8⫹) and bear the ␣␤ T-cell receptor (blue). Fetal animals generate more ␥␦ T cells than ␣␤ T cells, but the proportion of ␥␦ T cells generated drops off dramatically after birth. This early dominance of TCR␥␦ cells may have adaptive value: a large portion of these cells express nondiverse TCR specificities for common pathogen proteins and can mount a quick defense before the more traditional adaptive immune system has fully developed. For the rest of this chapter, we will focus on the development of TCR␣␤ T cells, which make up the vast majority of T lymphocytes in the body and are continually being generated, even after puberty, when the thymus shrinks considerably in size (involutes). DN Thymocytes Undergo ␤-Selection, Which Results in Proliferation and Differentiation Double-negative (DN) thymocytes that have successfully rearranged their TCR␤ chains are valuable, and are identified and expanded via a process known as ␤-selection (see Overview Figure 9-1). This process involves a protein that is uniquely expressed at this stage of development, a 33-kDa invariant glycoprotein known as the pre–T␣ chain. Pre-T␣ acts as a surrogate for the real TCR␣ chain, which has yet to rearrange, and assembles with a successfully rearranged and translated ␤ chain, as well as CD3 complex proteins. This precursor TCR/CD3 complex is known as the pre-TCR (Figure 9-4a) and acts as a sensor by initiating a signal transduction pathway. The signaling that the pre-TCR complex initiates is dependent on many of the same T-cell specific kinases used by a mature TCR (see Chapter 3), but does not appear to be dependent on ligand binding. In fact, little if any of the complex is expressed on the cell surface; rather, successful assembly of the complex may be sufficient to acti- CHAPTER 9 303 vate the signaling events. As we will learn in Chapter 10, B cells also express an immature receptor complex (the preBCR) that is coded by distinct genes, but is fully analogous to the pre-TCR. Pre-TCR signaling results in the following cascade of events: 1. 50 | 5. 6. Maturation to the DN4 stage (c-kitlow/⫺CD44⫺CD25⫺) Rapid proliferation in the subcapsular cortex Suppression of further rearrangement of TCR ␤-chain genes, resulting in allelic exclusion of the ␤-chain locus Development to the CD4⫹CD8⫹ double-positive (DP) stage Cessation of proliferation Initiation of TCR␣ chain rearrangement It is important to note that the proliferative phase prior to ␣ chain rearrangement enhances receptor diversity considerably by generating clones of cells with the same TCR ␤-chain rearrangement. Each of the cells within a clone can then rearrange a different ␣-chain gene, thereby generating an even more diverse population than if the original cell had undergone rearrangement at both the ␤- and ␣-chain loci prior to proliferation. TCR ␣-chain gene rearrangement does not begin until double-positive thymocytes stop proliferating. Most T cells fully rearrange and express a TCR␤ chain from only one of their two TCR alleles, a phenomenon known as allelic exclusion. Allelic exclusion is the result of inhibition of further rearrangement at the other TCR␤ allele (which must be fully rearranged to be expressed). This can be accomplished by reducing RAG expression so no more rearrangement can occur, as well as by making the locus inaccessible to further RAG interaction via more permanent changes in chromatin packaging. The details of the mechanisms responsible for this shutdown are still being investigated. However, negative feedback signals from a successfully assembled pre-TCR␤/pre-T␣ complex during ␤-selection clearly have a significant influence. Other events, including the proliferative burst that follows ␤-selection, which dilutes RAG protein levels, can also play a role. RAG levels continue to change after ␤-selection. They are restored after the proliferative burst and allow TCR␣ rearrangement to occur. They decrease once again after expression of a successfully assembled TCR␣␤ dimer.* * The mechanisms that turn off rearrangement at this (DP) stage of development are not as efficient as those that turn off additional TCR␤ chain rearrangement. In fact, both TCR␣ alleles successfully rearrange more frequently than we originally supposed. This means that many T cells actually express two TCR specificities: an interesting reality check that complicates a fundamentalist interpretation of the clonal selection hypothesis. 304 PA R T I V | Adaptive Immunity: Development (a) Pre-TCR (DN3, DN4, DP) (b) Mature αβTCR, immature signaling pathways (DP) TCR β Pre-Tα δ ε TCR α γ ε ζ ζ δ ε Signals γ ε ζ ζ Signals Cell becomes permissive for TCR α-chain locus rearrangement Stops additional TCR β-chain locus rearrangements (allelic exclusion) Stimulates expression of CD4 and CD8 coreceptors TCR β Stimulates proliferation Positive selection (intermediate affinity) and differentiation Negative selection (high-affinity and costimulatory signals) FIGURE 9-4 Changes in the structure and activity of the (c) Mature αβTCR, mature signaling pathways (SP) TCR α δ ε TCR β γ ε ζ ζ Signals Proliferation Effector function (cytokine secretion and/or cytotoxicity) Survival Once a young double-positive (DP) thymocyte successfully rearranges and expresses a TCR␣ chain, this chain will associate with the already produced TCR␤ chain, taking the place of the surrogate pre-TCR␣ chain, which is no longer actively expressed (Figure 9-4b). At this point, several “goals” have been accomplished by the thymus: hematopoietic cell precursors have expanded in the subcapsular cortex, committed to the T-cell lineage, and rearranged a set of TCR genes. They have also “chosen” to become a TCR␣␤ or TCR␥␦ T cell. This TCR␣␤ population now expresses both CD4 and CD8, and is ready for the second stage of T-cell development: selection. T-cell receptor through T-cell development. (a) The pre-TCR. The pre-TCR is assembled during the DN stage of development when a successfully rearranged TCR chain dimerizes with the nonvariant pre-T␣ chain. Like the mature ␣␤TCR dimer, the pre-TCR is noncovalently associated with the CD3 complex. Assembly of this complex results in intracellular signals that induce a variety of processes, including the maturation to the DP stage. (b) Mature ␣␤TCR expressed at the DP stage of development. Once a DP thymocyte has successfully rearranged a TCR␣ chain, it will dimerize with the TCR␤, replacing the preT␣ chain. This mature ␣␤TCR generates the signals that lead to either positive or negative selection (differentiation or death, respectively), depending on the affinity of the interaction. Note that the TCR␣ chain has a shorter intracellular region than the pre-T␣ chain and cannot generate intracellular signals independently. (c) Mature ␣␤TCR expressed at the SP stage. Although the ␣␤TCR/CD3 complex expressed by mature SP T cells is structurally the same as that expressed by DP thymocytes, the signals it generates are distinct. It responds to high-affinity engagement not by dying, but by initiating cell proliferation, activation, and the expression of effector functions. Low-affinity signals generate survival signals. The basis for the differences in signals generated by DP and SP TCR complexes is still unknown. Positive and Negative Selection CD4⫹CD8⫹ (DP) thymocytes, small, nonproliferating cells that reside in the thymic cortex, are the most abundant subpopulation in the thymus, comprising more than 80% of cells. Most important, they are the first subpopulation of thymocytes that express a fully mature surface TCR␣␤/CD3 complex and are therefore the primary targets of thymic selection (see Figure 9-1). Thymic selection shapes the TCR repertoire of DP thymocytes based on the affinity of their T-cell receptors for the MHC/peptides they encounter as they browse the thymic cortex. T-Cell Development Why is thymic selection necessary? The most distinctive property of the mature T cells is that they recognize only foreign antigen combined with self-MHC molecules. However, randomly generated TCRs will certainly have no inherent affinity for “foreign antigen plus self-MHC molecules.” They could just as well recognize foreign MHC/peptide combinations, which would not be useful, or self-MHC/self-peptide combinations, which could be dangerous. Two distinct selection processes are required: • Positive selection, which selects for those thymocytes bearing receptors capable of binding self-MHC molecules, resulting in MHC restriction • Negative selection, which selects against thymocytes bearing high-affinity receptors for self-MHC/peptide complexes, resulting in self-tolerance. Because only self-peptides are presented in the thymus, and only in association with self-MHC molecules, these two selection processes ensure that surviving thymocytes express TCRs that have low affinity for self-peptides in selfMHC. On the other hand, the processes do not guarantee that the T cells generated will bear receptors with high affinity for any specific self-MHC/foreign peptide combination. The ability of the immune system to respond to a foreign antigen depends on the probability that one of the millions of T cells that survives selection will bind one of the many MHC/peptide combinations expressed by an antigenpresenting cell that has processed pathogenic proteins outside the thymus. The vast majority of DP thymocytes (~98%) never meet the selection criteria and die by apoptosis within the thymus (Overview Figure 9-5). The bulk of DP thymocyte death (~95%) occurs among thymocytes that fail positive selection because their receptors do not specifically recognize self-MHC molecules. These cells do not receive survival signals through their TCRs, and die by a process known as death by neglect. A small percentage of cells (2%–5%) are eliminated by negative selection. Only 2% to 5% of DP thymocytes actually exit the thymus as mature T cells. Below, we briefly describe the experimental evidence for thymic involvement in MHC restriction. We also describe what is currently known about the positive and negative selection processes, as well as the models that have been advanced to explain what is known as “the thymic paradox.” Thymocytes “Learn” MHC Restriction in the Thymus Early evidence for the role of the thymus in selection of the T-cell repertoire came from chimeric mouse experiments performed by R. M. Zinkernagel and his colleagues. Recall that Zinkernagel won the Nobel Prize for showing that mature T cells are MHC restricted (i.e., they need to recog- | CHAPTER 9 305 nize antigen in the context of the MHC of their host to respond; see Classic Experiment in Chapter 8). However, he and his colleagues were curious to know how T cells became so “restricted.” They considered two possibilities: Either the T cell and APC simply had to have matching MHC types (i.e., an “A” strain T cell had to see an “A” strain antigenpresenting cell) or T cells, regardless of their own MHC type, “learned” the MHC type of their host sometime during development. Zinkernagel and colleagues thought that such learning could take place in the thymus, the T-cell nursery. To determine if T cells could be “taught” to recognize the host MHC, they removed the thymus (thymectomized) and irradiated (A  B) F1 mice so they had no functional immune system (Figure 9-6). They then reconstituted the hematopoietic cells with an intravenous infusion of F1 bone marrow cells, but replaced the thymus with one from a B-type mouse. (To be certain that the thymus graft did not contain any mature T cells, they irradiated it before transplantation.) In this experimental system, T-cell progenitors from the (A  B) F1 bone marrow would mature within a thymus that expresses only B-haplotype MHC molecules on its stromal cells. Would these (A  B) F1 T cells now be MHC restricted to the B haplotype of the thymus in which they developed? Or, because they expressed both A and B MHC, would they be able to recognize both A and B MHC haplotypes? To answer this question, the investigators infected the chimeric mice with lymphocytic choriomeningitis virus (LCMV, the antigen) and removed the immunized, mature splenic T cells to see which LCMV-infected target cells (APCs) they could kill. They tested them against infected APCs from strain A, strain B, and strain A  B mice. As shown in Figure 9-6, T cells from the chimeric mice could only lyse LCMVinfected target cells from strain B mice. Thus, the MHC haplotype of the thymus in which T cells develop determines their MHC restriction. T cells “learned” which MHC haplotype they are restricted to during their early days in the thymus. Although once we referred to this process as “T-cell education,” we now know that it is the consequence of a brutal selection process. T Cells Undergo Positive and Negative Selection In the thymus, thymocytes come into contact with thymic epithelial cells that express high levels of class I and class II MHC molecules on their surface. These self-MHC molecules present self-peptides, which are typically derived from intracellular or extracellular proteins that are degraded in the normal course of cellular metabolism. DP thymocytes undergo positive and negative selection, depending on the signals they receive when they encounter self-MHC/selfpeptide combinations with their TCRs. While working to understand negative and positive selection, it helps to picture an early model that was advanced by several seminal immunologists (see Overview 9-5 OVERVIEW FIGURE Positive and Negative Selection of Thymocytes in the Thymus Cortex CD8 CD3 T-cell receptor DP thymocyte 2%–5% high-affinity interaction Negative selection cTEC CD4 90–96% no interaction 2%–5% low/int-affinity interaction Death by neglect Positive selection Class I and/or Class II MHC molecules CD4+ TH cell CD8+ TC cell Medulla Negative selection AIRE+ CD4+ CD8+ mTEC TH cell TC cell Dendritic cell Mature CD4+ or CD8+ T lymphocytes emigrate to periphery Thymic selection involves multiple interactions of DP and SP thymocytes with both the cortical and medullary thymic stromal cells, as well as dendritic cells and macrophages. Selection results in a mature T-cell population that is both self-MHC restricted and self-tolerant. DP thymocytes that express new TCR␣␤ dimers browse the MHC/ peptide complexes expressed by the cortical thymic epithelial cells (cTECs). The large majority of DP thymocytes die in the cortex by neglect because of their failure to bind MHC/peptide combinations with sufficient affinity. The small percentage whose TCRs bind MHC/ 306 peptide with high affinity die by clonal deletion (negative selection). Those DP thymocytes whose receptors bind to MHC/peptide with intermediate affinity are positively selected and mature to singlepositive (CD4⫹ or CD8⫹) T lymphocytes. These migrate to the medulla, where they are exposed to AIRE⫹ medullary thymic epithelial cells (mTECs), which express tissue-specific antigens and can mediate negative selection. Medullary dendritic cells can acquire mTEC antigens by engulfing mTECs, and mediate negative selection (particularly of MHC Class II restricted CD4⫹ thymocytes). T-Cell Development CONTROL Infect with LCMV (A × B)F1 Spleen cells LCMV-infected strain A cells Killing LCMV-infected strain B cells Killing EXPERIMENT (A × B) F1 (H–2a/b) 1 Thymectomy 2 Lethal x-irradiation Strain B thymus graft (H–2b) (A × B) F1 hematopoietic stem cells (H–2a/b) Infect with LCMV Spleen cells LCMV-infected strain A cells No killing LCMV-infected strain B cells Killing FIGURE 9-6 Experimental demonstration that the thymus selects for maturation only those T cells whose T-cell receptors recognize antigen presented on target cells with the haplotype of the thymus. Control (A ⫻ B) F1 mice infected with LCMV produce mature T cells that are able to kill both infected-strain A cells and infected-strain B cells, demonstrating that these T cells are restricted to MHC molecules expressed by both strain A (H-2a) and strain B (H-2b). In order to determine the involvement of the thymus in the restriction specificity of T cells, investigators grafted thymectomized and lethally irradiated (A ⫻ B) F1 (H-2a/b) mice with a strain-B (H-2b) thymus and reconstituted it with (A ⫻ B) F1 bone marrow cells. After infection with LCMV, the CD8⫹ cytotoxic (CTL) cells from this reconstituted mouse were assayed for their ability to kill 51Cr-labeled strain A or strain B target cells infected with LCMV. Only strain B target cells were lysed, suggesting that the H-2b grafted thymus had selected for maturation only those T cells that could recognize antigen combined with H-2b MHC molecules. | CHAPTER 9 307 Figure 9-5). Although we now know that some aspects of this model are too simplistic, its core principles remain relevant and provide a very useful framework for understanding thymic selection. Briefly, DP thymocytes can be considered to have one of three fates, depending on the affinity of their new T-cell receptors for the MHC/selfpeptide combinations that they encounter: If their newly generated TCRs do not bind to any of the MHC/self-peptides they encounter on stromal cells, they will die by neglect. If they bind too strongly to MHC/self-peptide complexes they encounter, they will be negatively selected. If they bind with a low, “just right” affinity to MHC/self-peptide complexes, they will be positively selected and mature to the singlepositive stage of development. (We will also see that cells that are positively selected in the cortex undergo a second, very important round of negative selection in the medulla, following lineage commitment to CD4⫹ or CD8⫹ cells.) We review the origins of the model and discuss some of the modifications that have been suggested by recent data below. See Classic Experiment Box 9-1 for experimental evidence in support of this model from the earliest TCR transgenic mouse. Thymic stromal cells, including epithelial cells, macrophages, and dendritic cells, play essential roles in positive and negative selection and should be part of our visualization of thymic selection events. Young DP thymocytes are in intimate contact with these cells and “browse” the MHC/ self-peptides displayed on their surfaces. Each of these cell types has the capacity to express high levels of class I and class II MHC proteins, and typically have extended processes that contact many developing thymocytes. Some also express costimulatory ligands, including CD80 (B7-1) and CD86 (B7-2). As the thymocytes migrate through the thymus, they encounter multiple different stromal cell surfaces, and have the opportunity to bind many different MHC/ peptide combinations. Positive Selection Ensures MHC Restriction If a CD4⫹CD8⫹ thymocyte recognizes a self-MHC/peptide complex on the cortical epithelial cells that they browse, it will undergo positive selection, a process that induces both survival and differentiation of DP thymocytes. Remarkably, the majority of newly generated thymocytes do not successfully engage the MHC/peptides they encounter with their TCRs. Either they have not generated a functional TCR␣␤ combination or the combination does not bind MHC/peptide complexes with sufficient affinity. These cells have “failed” the positive selection test and die by apoptosis within 3 to 4 days. TCR/MHC/self-peptide interactions that initiate positive selection are thought to be at least three times lower in intensity than interactions that initiate negative selecting signals (Figure 9-7). The signal cascade generated by positive selecting TCR interactions has not yet been fully characterized. We do know that many of the signaling molecules required for CLASSIC EXPERIMENT Insights about Thymic Selection from the First TCR Transgenic Mouse Have Stood the Test of Time In the late 1980s Harald von Boehmer and colleagues published a series of seminal experiments that provided direct and compelling evidence for the influence of T-cell receptor interactions on positive and negative selection. They recognized that in order to understand how positive and negative selection worked, they would need to be able to trace and compare the fate of thymocytes with defined T-cell receptor specificities. But how could one do that if every one of the hundreds of millions of normal thymocytes expresses a different receptor? The researchers decided to develop a system in which all thymocytes expressed a T-cell receptor of known specificity, and thereby generated one of the first TCR transgenic mice. Transgenic animals are made by injecting a gene under the control of a defined promoter into a fertilized egg (zygote) (see Chapter 20). The gene—in fact, often many copies of the gene—will be incorporated randomly into the genome of the zygote. Therefore, the gene will be present in all cells of the mouse that develop from that zygote; however, only those cells that can activate the promoter will express the gene. The team of von Boehmer et al. developed the first TCR transgenic animals by isolating fully rearranged TCR␣ and TCR genes of a mature CD8⫹ T-cell line that was known to be specific for an antigen expressed only in male mice (the “H-Y” antigen) in the context of H-2Db Class I MHC molecules (Figure 1). They generated a genetic construct that included both rearranged genes under the control of a T-cell specific promoter and injected this into a mouse zygote. Since the receptor transgenes were already rearranged, other TCR gene rearrangements were suppressed in the transgenic mice (the phenomenon of allelic exclusion; see text and Chapter 7); therefore, a very high percentage of the thymocytes in the transgenic mice expressed the TCR␣␤ combination specific for the H-Y/H-2Db. (It turns out that not all thymocytes expressed this TCR combination. Why not? 308 Recall that allelic exclusion does not operate well for the TCR␣ locus in particular.) The team’s choice of specificities was very clever. Because all thymocytes expressed an anti-male antigen TCR specificity, the researchers could directly compare the phenotype of autoreactive thymocytes and non-autoreactive thymocytes simply by comparing male and female mice in the same litter. Because the MHC restriction of the TCR was known, they could also observe the influence of MHC haplotype on thymocyte development simply by breeding the mice to other inbred strains. The results of a comparison of CD4⫹ versus CD8⫹ phenotype among thymocytes in H-2Db male versus female mice is shown in Figure 1. The primary data from one of the first publications are shown in CTL H-Y specific H-2Db restricted × H-Y peptide Clone TCR α and β genes Male cell (H-2Db ) α Female cell (H-2Db ) β Use to make α H-Y TCR transgenic mice Male H-2Db Female H-2Db Female H-2Dd + − − CD4− 8− ++ + + CD4+ 8+ + ++ ++ CD4+ + + − CD8+ − ++ − H-Y expression Thymocytes FIGURE 1 Experimental demonstration that negative selection of thymocytes requires both self-antigen and self-MHC, and positive selection requires self-MHC. In this experiment, H-2Db male and female transgenics were generated by injecting TCR transgenes specific for H-Y antigen plus the Db MHC molecule into zygotes. The H-Y antigen is expressed only in males. H-2Dd mice were also generated by backcrossing these transgenics onto an H-2d strain (e.g., BALB/c). FACS analysis of thymocytes from the transgenics showed that mature CD8⫹ T cells expressing the transgene were absent in the male H-2Db mice but present in the female H-2Db mice, suggesting that thymocytes that bind with high affinity to a self antigen (in this case, H-Y antigen in the male mice) are deleted during thymic development. Results also show that DP but not CD8⫹ thymocytes develop in female H-2Dd mice, which do not express the proper self-MHC. These findings indicate that positive selection and maturation require self-MHC interactions. [Adapted from H. von Boehmer and P. Kisielow, 1990, Science 248:1369.] BOX 9-1 Figure 2 and depict flow cytometric data of CD4 and CD8 expression of thymocytes. The flow cytometric profile (see Chapter 20) of a typical wild-type thymus is shown in Figure 2a and can be used for comparison. In this profile, the CD8-expression status of a cell is shown on the X-axis, the CD4expression status on the Y-axis. The data are quantified in the upper right corner of the profile. As you can see, more than 80% of normal thymocytes are DP, 10% to 15% are CD4 single positive, 3% to 5% are CD8 single-positive cells, and only a small percentage of cells are DN. How do the CD4 versus CD8 phenotypes of male and female H-Y TCR transgenic mice (Figure 2b) differ, and what does this tell you? Thymocytes in female mice are in all stages of development: DN, DP, and SP. In fact, they seem to have an abundance of mature CD8⫹ SP thymocytes. We’ll come back to this below. The male mice, however, have few if any anti-H-Y TCR transgenic DP thymocytes. These data show that in an environment where DP thymocytes are exposed to the antigen for which they are specific (in this case the male H-Y antigen), they are eliminated. This result was fully consistent with the concept of negative selection and showed that self-reactive cells were removed from the developing repertoire, perhaps at the DP stage. However, the mice offered insights into many more aspects of thymic selection. Evidence for the role of TCR interactions in positive selection, for instance, came (a) Wild-type thymus 108 cells (b) from a different experiment in which the investigators asked what would happen if the “correct” restricting element, H-2Db, was not present in the mouse. To do this they performed backcrosses of their TCR transgenic mice to mice that expressed a different H-2 haplotype, an H-2Dd mouse. (For those who want to brush up on their genetics, determine what crosses you would make to do this, and include a plan for assessing the genotype of your mice.) Figure 2c shows the CD4 versus CD8 phenotype of H-Y TCR transgenic thymocytes from female H-2Db/d mice and H-2Dd mice. What did they find? H-2Dd females had no mature T cells, indicating that in the absence of the MHC haplotype to which the TCR was restricted, thymocytes could not mature beyond the DP stage. These data provided direct evidence for the necessity of a TCR/self-MHC interaction for thymocyte maturation (positive selection) to occur. These “simple” experiments also revealed a third feature of thymic development and initiated a controversy that continues to this day. Return to Figure 2b and take a look at the female mouse data. Unlike wild-type thymocytes, which typically include 3%- to -5% CD8⫹ T cells, 46% of developing thymocytes in the anti-H-Y TCR transgenic were mature CD8⫹ cells. What did this mean? Recall that the T-cell line that the TCR genes came from originally was a CD8⫹ T-cell line, restricted to Class I MHC (H-2Db). These data showed that the MHC restriction specificity Male H-2Db Female H-2Db (c) of the TCR (in this case, a restriction for Class I) influenced its CD4 versus CD8 lineage choice. In other words, the results suggested that thymocytes with new TCRs that preferentially bound Class II would become CD4⫹ SP mature T cells, and thymocytes with TCRs that preferentially bound Class I would develop into CD8⫹ SP mature T cells. The fundamental conclusions of this now classic set of experiments have stood the test of time and have been supported by many different, equally clever experiments: (1) Negative selection of autoreactive cells results in their elimination at the DP stage and beyond. (It is important to note that it later became clear that the promoter driving these first H-Y TCR transgenes permitted earlier-than-normal expression of the TCR, which resulted in elimination of autoreactive cells at an earlier-than-usual point in development. Experiments now indicate that negative selection by clonal deletion typically occurs after positive selection as cells transit between the DP and SP stages of development.) (2) Positive selection involves a low affinity interaction between thymocyte TCR and self-MHC/ peptide interactions. (3) Commitment to the CD4 versus CD8 lineage is determined by the preference of the TCR for MHC Class I versus Class II. How this last event happens remains a topic of intense controversy (see text). von Boehmer, H., H. Teh, and P. Kisielow. 1989. The thymus selects the useful, neglects the useless and destroys the harmful. Immunology Today 10:57–61. Female H-2Db/d Female H-2Dd 11 83 3 3 ⫹ CD4 ⫺ ⫺ CD8 ⫹ FIGURE 2 Primary data from experiments summarized in Figure 1. The relative staining by an anti-CD8 antibody is shown on the X-axis, and the relative staining by an anti-CD4 antibody appears on the Y-axis. DP, DN, CD4 SP, and CD8 SP phenotypes are divided into quadrants, and percentages are given by quadrant. (a) CD4 versus CD8 phenotype of a wild-type thymus. (b) CD4 versus CD8 phenotype of thymocytes from H-Y TCR transgenic H-2b male and female. (c) CD4 versus CD8 phenotype of thymocytes from H-Y TCR transgenic H-2b/d and H-2d/d females. [Parts (b) and (c) from H. von Boehmer and P. Kisielow, 1990, Science 248:1369.] 309 PA R T I V 310 | Adaptive Immunity: Development Cell number Negative Selection (Central Tolerance) Ensures Self-Tolerance TCR affinity for self Affinity too low Intermediate affinity Affinity too high Fail to be positively selected Positively selected Negatively selected “Death by neglect” Survive Deleted 90%–96% 2%–5% 2%–5% FIGURE 9-7 Relationship between TCR affinity and selection. This figure schematically illustrates the association between thymocytes’ fate and the affinity of their TCR for self-MHC/peptide complexes that they encounter in the thymus. Fewer than 5% of thymocytes produce TCRs that bind to MHC/peptide complexes with high affinity. Most of these will be deleted by negative selection (some will become regulatory T cells and other specialized cell types). More than 90% generate TCRs that either do not bind to MHC/ peptide complexes or bind them with very low affinity. These die by neglect. Fewer than 5% generate TCRs that bind with just the right intermediate affinity to self-MHC/peptide complexes. These will survive and mature. positive selection are also involved in T-cell receptor mediated activation of mature T cells (see Chapter 3), an event, however, that requires higher-affinity TCR interactions. We still do not fully understand how low-affinity positive selecting signals initiate maturation to the single-positive stage, or how they differ from pro-apoptotic negative selecting signals, which are also TCR mediated. What is the adaptive value of positive selection? Why didn’t we evolve a system that negatively selects but leaves the T-cell receptor repertoire otherwise intact? Some have suggested that the paring down of the repertoire is important to increasing the efficiency of negative selection, as well as to increasing the efficiency of peripheral T-cell responses. Without positive selection, the system would be cluttered with a great many cells that are unlikely to recognize anything, and reduce the probability that a T cell will find “its” antigen in a reasonable period of time. This is a reasonable speculation; however, positive selection may also offer more subtle advantages that have often inspired discussion in the field. Autoreactive CD4⫹CD8⫹ thymocytes with high-affinity receptors for self-MHC/self-peptide combinations are potentially dangerous to an organism, and many are killed by negative selection in the thymus. In fact, errors in the negative selection process are responsible for a host of autoimmune disorders, including Type 1 diabetes (see Clinical Focus). Negative selection is defined broadly as any process that rids a repertoire of autoreactive clones and is responsible for central tolerance. It is likely that most negative selection occurs via a process known as clonal deletion, where high-affinity TCR interactions directly induce apoptotic signals. Clonal deletion of DP thymocytes appears to be optimally mediated by the same cells (APCs) and same interactions (high-affinity TCR engagement coupled with costimulatory signals) that activate mature T cells. Why strong TCR signals result in death of immature T cells, but proliferation and differentiation of mature T cells (see Figure 9-4c) remains an active area of investigation. Thymic dendritic cells and macrophages, which are found in multiple areas of the thymus, clearly have the ideal features to mediate negative selection, but interestingly, so do medullary epithelial cells, which express high levels of the costimulatory ligands CD80 and CD86 as well as a unique transcription factor that allows them to present tissue-specific antigens (see below). Therefore, both the cortex and the medulla have the potential to induce negative selection. Do We Delete Thymocytes Reactive to Tissue-Specific Antigens? We have seen that MHC molecules can present peptides derived from both endogenous and exogenous proteins on the surface of antigen-presenting cells (see Chapter 8). However, only a fraction of cell types—thymocytes, stromal cells, macrophages, and other antigen-presenting cells— reside in the thymus, so we would expect these cells to produce only a subset of the proteins encoded in the genome. How, then, can the thymus possibly get rid of developing T cells that are autoreactive to tissue-specific antigens—for example, proteins specific to the brain, the liver, the kidney, or the pancreas? This question bothered immunologists for a long time. For a bit, we assumed that other mechanisms of tolerance in the periphery took care of autoreactivity to tissue-specific proteins, but investigations in the late 1990s surprised us all and revealed that the thymus had an extraordinary capacity to express and present proteins from all over the body. Bruno Kyewski and colleagues showed that this capacity was a unique feature of thymic medullary epithelial cells. Diane Mathis, Cristoph Benoist, and their colleagues went on to show that some medullary epithelial cells express a unique protein, AIRE, that allows cells T-Cell Development | CHAPTER 9 311 BOX 9-2 CLINICAL FOCUS How Do T Cells That Cause Type 1 Diabetes Escape Negative Selection? We have an extensive appreciation of the pain caused by autoimmune disease and its clinical progression, but although we have also gained a deeper understanding of the mechanisms behind immune tolerance, we still know little about the precise origins of autoimmune disorders. Indisputably, the primary cause of autoimmune disease is the escape of self-reactive lymphocytes—B cells, T cells, or both—from negative selection. Surprisingly little is known about the reasons for this escape, or the specificity and phenotype of the escapees that cause even the best-characterized disorders, including Type I diabetes (T1D), multiple sclerosis (MS), rheumatoid arthritis (RA), and systemic lupus erythematosus (SLE). Recent work on a mouse strain, the non-obese diabetic or NOD mouse, which is markedly susceptible to Type 1 diabetes, has shed light on the features of the self-reactive T cells that cause the disease, and revealed some interesting reasons for their escape from central tolerance. T1D is a T-cell mediated autoimmune disease caused by T cells that kill insulin-producing beta islet cells in the pancreas. Many of the autoreactive T cells actually recognize a specific peptide from the insulin protein itself. When Emil Unanue and colleagues examined the fine specificity of T cells that had infiltrated the islet cells in diabetesprone mice, they found two types of cells: T cells that could respond to the insulin peptide that was generated via intracellular MHC Class II processing pathways by dendritic cells (“Type A” T cells), and an unexpected population of T cells that responded to the insulin peptide only if it were added exogenously to the MHC Class II molecules on dendritic cells (“Type B” T cells). The investigators speculate that the endogenously processed MHC/ insulin peptide combination and the exogenously formed MHC/insulin peptide complex differ in conformation and therefore activate different T-cell receptor specificities. Perhaps more important, these observations suggest an elegant and precise reason for the escape of autoreactive T cells from the thymus. Dendritic cells in the pancreas, but not other tissues, appear uniquely capable of forming MHC Class II/insulin peptide from exogenous sources, presumably because they are in an environment replete with extracellular insulin secreted by the beta islet cells. Thymic medullary epithelial cells or dendritic cells would not have this capacity, and therefore “Type B” T cells would sneak to express, process, and present proteins that are ordinarily only found in specific organs. This group discovered AIRE while studying the molecular basis for autoimmunity by examining human disease. They traced a mutation that caused autoimmune polyendocrinopathy syndrome 1 (APS1, otherwise known as APECED, see Chapter 18) to this gene, and called it AIRE for “autoimmune regulator.” The title of the paper that announced their discovery, “Projection of an Immunological Self Shadow within the Thymus by the AIRE Protein,” aptly described AIRE’s powerful function. AIRE’s mechanism of action is still under investigation. It has classic features of a transcription factor and may be part through the negative selection process in the thymus. Why “Type A” T cells also escape is less clear. They are autoreactive to more conventionally processed insulin/MHC Class II complexes, which are clearly present on thymic medullary cells. Perhaps Type A cells expressed too low an affinity for insulin/MHC Class II peptides in the thymus, but were inspired by the high levels of insulin and inflammatory signals in a diabetic islet? Perhaps they escaped negative selection because the level of insulin peptide expression on medullary epithelium was too low in the NOD mouse strain? In fact, recent data from both mice and humans suggest that the level of insulin expression significantly influences the progression of disease and the efficiency of negative selection in the thymus. Understandably, most current therapies for autoimmune disease focus on inhibiting the secondary but most proximal cause of autoimmune disease: the peripheral activation of autoreactive T and B cells that escaped negative selection. However, by defining the molecular reasons for self-reactive lymphocyte escape, we may ultimately find a way to develop therapeutic approaches to correct negative selection defects, too. of a transcriptional complex that facilitates expression of tissue-specific genes by regulating not only translation, but also chromatin packing. Thus, it allows medullary epithelial cells to express proteins not ordinarily found in the thymus, process them, and present them in MHC molecules. This is particularly useful for the presentation of MHC Class I peptides and negative selection of CD8⫹ thymocytes. However, neighboring dendritic cells and macrophages are also thought to phagocytose medullary epithelial cells and so can then present their protein contents in MHC Class II molecules, and mediate negative selection of CD4⫹ thymocytes (see Overview Figure 9-5). 312 PA R T I V | Adaptive Immunity: Development Other Mechanisms of Central Tolerance Other mechanisms of thymic negative selection (central tolerance) that do not necessarily involve cell death have been proposed. They include clonal arrest, where thymocytes that express autoreactive T-cell receptors are prevented from maturation; clonal anergy, where autoreactive cells are inactivated, rather than deleted; and clonal editing, where autoreactive cells are given a second or third chance to rearrange a TCR␣ gene. Each of these mechanisms has some experimental support, but clonal deletion is probably the most common mechanism responsible for thymic negative selection. The generation of regulatory T cells from autoreactive thymocytes can certainly be considered of importance among central tolerance mechanisms and will be discussed below. Superantigens Superantigens, viral or bacterial proteins that bind simultaneously to the V domain of a T-cell receptor and to the ␣ chain of a Class II MHC molecule (outside the peptide binding groove), can induce activation of all T cells that express that particular family of V chains (see Figure 11-5). Superantigens are also expressed in the thymus of mice and humans and influence thymocyte maturation. Because they mimic high-affinity TCR interactions, superantigens will induce the negative selection of all DP thymocytes whose receptors possess V domains targeted by the superantigen. However, because we continually generate T cells with a wide range of T-cell receptor specificities, this loss does not appear to have major clinical consequences. The Selection Paradox: Why Don’t We Delete All Cells We Positively Select? Full understanding of the process of positive and negative selection requires an appreciation of the following paradox: If positive selection allows only thymocytes reactive with self-MHC molecules to survive, and negative selection eliminates self-MHC-reactive thymocytes, then why are any T cells allowed to mature? Other factors must operate to prevent these two MHC-dependent processes from eliminating the entire repertoire of MHC-restricted T cells. The most straightforward model advanced to explain the paradox of MHC-dependent positive and negative selection is the affinity hypothesis,* which asserts that differences in the strength of TCR-mediated signals received by thymo- * Note that a similar version of this hypothesis is also sometimes referred to as the avidity hypothesis. Avidity and affinity do have distinct meanings, which can differ for different investigators. For the sake of simplicity, we consider affinity to mean the strength of interaction between the TCR and its MHC/peptide ligand. For a more nuanced discussion of the differences between affinity and avidity and their influence on thymic development, see the review by Kyewski listed in the reference section of this chapter. cytes undergoing positive and negative selection determine the outcome of the interaction. Double-positive thymocytes that receive low-affinity signals would be positively selected, those that received high-affinity (autoreactive) signals would be negatively selected, and those that received no signal at all would die by neglect (see Figure 9-7 and Overview Figure 9-5). The OT-I TCR transgenic mouse, developed by Kristin Hogquist and colleagues, is a superb model for the study of thymic selection using TCRs and MHC/peptides with known affinities. Not only is its peptide/MHC specificity known precisely (it binds a chicken ovalbumin peptide in the context of the H-2Kb MHC Class I molecule), but its affinity for a range of peptides that vary in sequence has also been defined. To determine the influence of affinity on T-cell development, investigators also took advantage of two other immunological tools: (1) the fetal thymic organ culture (FTOC) system, where thymic development can be followed in vitro (Figure 9-8a), and (2) mice defective in MHC Class I antigen processing. Several different mutations that lead to defective MHC Class I presentation exist. An experimental scheme using the TAP-1 knock out (TAP-1⫺), a mutant mouse where newly generated peptides cannot access newly generated MHC Class I molecules (see Chapter 8), is depicted in Figure 9-8b. These mice cannot load their MHC Class I with endogenous peptides; they are expressed on the cell surface but are “empty” and do not hold their shape. However, they can be loaded with exogenous peptides, which stabilize their conformation. Therefore, by incubating fetal thymic organs from these mutant mice with peptides of choice, investigators were able to control the type of peptide presented by the MHC Class I. What did investigators find? As expected, OT-I⫹/TAP-1⫹ fetal thymic organ cultures produced more CD8⫹ cells than wild-type FTOC. This is because virtually all of their thymocytes expressed a receptor that bound to the variety of MHC Class I (H-2Kb)/self-peptide complexes expressed by the normal stroma with intermediate affinity, generating a signal that results in positive selection (see Classic Experiment Box 9-1). Also, as expected, CD8⫹ mature T cells failed to develop in OT-I⫹/TAP-1⫺ thymic organ cultures because there was no normal MHC Class I and, therefore, no TCR signal—a failure of positive selection, resulting in death by neglect (see Figure 9-8b, top row). However, when a low-affinity peptide was added to the OT-I⫹/TAP-1⫺ culture, MHC Class I was able to load itself with peptide and assume a normal conformation (see Figure 9-8b, middle row). Positive selection occurred, and CD8⫹ T cells developed successfully. Addition of high-affinity peptides induced a strong TCR signal and negative selection, resulting in deletion of DP cells in OT-I⫹/TAP-1⫺ thymic organ cultures (see Figure 9-8b, bottom row). Interestingly, low concentrations of high-affinity peptides also induced positive selection, although the cells that developed were not fully functional. These results and many others provided clear support for the affinity model, showing that the T-Cell Development | CHAPTER 9 313 (a) Experimental procedure—fetal thymic organ culture (FTOC) Remove thymus Place in FTOC Porous membrane Growth medium (b) Peptide added Signal generated CD8+ T cell selection None cTEC None None Empty MHC class I Intermediate Low affinity cTEC Positive MHC class I filled with lowaffinity peptide Strong High affinity cTEC Negative MHC class I filled with highaffinity peptide FIGURE 9-8 Role of TCR affinity for peptide in thymic selection. Fetal thymocyte populations have not yet undergone positive and negative selection and can be easily manipulated to study the development and selection of single-positive (CD4⫹CD8⫺ and CD4⫺CD8⫹) T cells. (a) Outline of the experimental procedure for in vitro fetal thymic organ culture (FTOC). Fetal thymi are cultured on a filter disc at the interface between medium and air. Reagents added to the medium are absorbed by the thymi. In this experiment, peptide is added to the medium of FTOC from TAP-1 knockout (TAP-1⫺) mice, which are unable to form peptide-MHC Class I complexes unless peptide is added exogenously to the culture medium. (b) The development and selection of CD8⫹CD4⫺ class I–restricted T cells strength of TCR/MHC/peptide interaction does, indeed, influence cell fate. The OT-I system has also been used to estimate the range of TCR affinities that define positive versus negative selection outcomes. It appears as if negative selection occurs at affinities that are threefold higher than those that induce positive selection. depends on TCR peptide–MHC class I interactions. The mice used in this study were transgenic for OT-I TCR ␣ and  chains, which recognize an ovalbumin peptide when associated with MHC Class I; the proportion of CD8⫹ T cells that develop in these thymi is higher than normal because all thymocytes in the mouse express the MHC Class I restricted specificity. Different peptide/MHC complexes interact with the TCR with different affinities. Varying the peptide added to FTOC from OT-I⫹/TAP-1⫺ mice revealed that peptides that interact with low affinities (middle row) resulted in positive selection, and levels of CD4⫺CD8⫹ cells that approached that seen in the OT-I⫹/ TAP-1⫹ mice. Peptides that interact with high affinities (bottom row) cause negative selection, and fewer CD4⫺CD8⫹ T cells appeared. An Alternative Model Can Explain the Thymic Selection Paradox Philippa Marrack, John Kappler, and their colleagues thoughtfully offered an alternative possibility to explain the “thymic paradox”—that is, why we don’t negatively select all cells that we positively select. These investigators advanced 314 PA R T I V | Adaptive Immunity: Development the altered peptide model, a suggestion that cortical epithelium, which induces positive selection, makes peptides that are unique and distinct from peptides made by all other cells, including the thymic cells that induce negative selection. Thus, those thymocytes selected on these unique peptide/ MHC complexes would not be negatively selected when they browse the medulla and other negatively selecting cells. Initial experiments did not support this model; however, advances in our ability to analyze peptides presented on cell surfaces in detail suggests the model has merit. Cortical epithelial cells may, indeed, process peptides differently, and are likely to present a different array of peptides to developing T cells. Recall that peptides are processed intracellularly by several mechanisms, including the activity of the proteosome (see Figure 8-16). It appears that components of the proteosome expressed by cortical epithelial cells (the “thymoproteasome”) are unique. The thymic epithelium may also express unique versions of proteases (e.g., cathepsins). The altered peptide and affinity models are by no means mutually exclusive. Investigations firmly establish that affinity plays an important role in distinguishing positive from negative selection. That the cortex may also generate novel peptides simply underscores the possibility that the thymus evolved multiple ways to ensure that we would be able to generate a sufficiently large pool of T cells with diverse specificities. Do Positive and Negative Selection Occur at the Same Stage of Development, or in Sequence? Strictly interpreted, the affinity model assumes that positive and negative selection operate on exactly the same target population (DP thymocytes) and in the same microenvironment of the thymus (the thymic cortex). Although it is clear that thymocytes undergo positive selection in the cortex, as we have seen above, the medulla is the site of negative selection to tissue-specific antigens. This compartmentalization of function suggests some thymocytes are positively selected and initiate a maturation program prior to negative selection. That this can happen is supported by observations showing that “semi-mature” thymocytes—those that have been positively selected and are in transition from the DP to the SP stages—are excellent targets of negative selection. Most likely, thymocytes can be negatively selected at more than one point of development (illustrated in Overview Figure 9-5). Those thymocytes that bind with high affinity as they browse MHC/peptide complexes in the cortex are negatively selected, and those that receive positively selecting signals are given permission not only to mature, but to migrate to the cortical medullary boundary and ultimately to the medulla. In the medulla, semi-mature thymocytes can again be subject to negative selection if they interact with high affinity to tissue-specific antigens. Migration of positively selected thymocytes to the medulla is now known to be dependent on expression of the CCR7 chemokine receptor: CCR7 deficient thymocytes fail to enter the medulla. Interestingly, they still mature and are exported to the periphery, but there they cause autoimmunity— an observation that, again, underscores the central importance of the medulla in removing autoreactive cells from the T-cell repertoire. Lineage Commitment As thymocytes are being screened on the basis of their TCR affinity for self-antigens, they are also being guided in their lineage decisions. Specifically, a positively selected doublepositive thymocyte must decide whether to join the CD8⫹ cytotoxic T-cell lineage or the CD4⫹ helper T-cell lineage. Lineage commitment requires changes in genomic organization and gene expression that result in (1) silencing of one coreceptor gene (CD4 or CD8) as well as (2) expression of genes associated with a specific lineage function. Immunologists continue to debate about the cellular and molecular mechanisms responsible for lineage commitment. We will review the most current perspectives. Several Models Have Been Proposed to Explain Lineage Commitment When early studies with TCR transgenic mice revealed that affinity for MHC Class I versus Class II preference dictated the CD8⫹ versus CD4⫹ fate of developing thymocytes (see Classic Experiment Box 9-1, investigators advanced two simple testable models to explain how developing T cells “matched” their TCR preference with their coreceptor expression. In the instructive model, TCR/CD4 and TCR/CD8 coengagement generates unique signals that directly initiate distinct developmental programs (Figure 9-9a). For example, if a thymocyte randomly generated a TCR with an affinity for MHC Class I, the TCR and CD8 would bind MHC Class I together, and generate a signal that specifically initiated a program that silenced CD4 expression and induced expression of genes specific for cytotoxic T-cell lineage function. Likewise, TCR/CD4 coengagement would generate a unique signal that initiated CD8 silencing and the helper T-cell developmental program. In the stochastic model, a positively selected thymocyte randomly down-regulates CD4 or CD8 (Figure 9-9b). Only those cells that express the “correct” coreceptor—the ones that can coengage MHC with the TCR—generate a TCR signal strong enough to survive to mature. In this model, TCR/CD4 and TCR/CD8 coengagement does not necessarily generate distinct signals. Unfortunately, studies that followed the consequences of such mismatches confounded researchers by providing evidence in support of both models! Clearly they were too simplistic. Further experiments challenged some of the assumptions on which these early efforts were based and indicated that T-Cell Development (a) Instructive model CD4+ 8+ | CHAPTER 9 315 CD8 engagement signal CD4lo 8hi CD4− 8+ T cell CD4 engagement signal CD4+ 8+ CD4hi 8lo CD4+ 8− T cell (b) Stochastic model Able to bind Ag + class I MHC CD4+ 8+ Random CD4 CD4− 8 + T cell Not able to bind Ag + class I MHC CD4lo 8hi Apoptosis Able to bind Ag + class II MHC CD4+ 8+ Random CD8 CD4+ 8 − T cell Not able to bind Ag + class II MHC CD4hi 8lo Apoptosis (c) Kinetic signaling model Disrupted TCR signal, IL–7 CD4–8+ T cell Engagement signal CD8 CD4+8+ CD4+8lo Continuous TCR signal CD4+8– T cell FIGURE 9-9 Proposed models of lineage commitment, the decision of double-positive thymocytes to become helper CD4ⴙ or cytotoxic CD8ⴙ T cells. (a) According to the instructive model, interaction of a coreceptor with the MHC molecule for which it is specific results in down-regulation of the other coreceptor. (b) According to the stochastic model, down-regulation of CD4 or CD8 is a random process. (c) According to the kinetic signaling model, the decision to commit to the CD4 or CD8 lineage is based on the continuity of the TCR signal that a thymocyte receives. Positive selection results in down-regulation of CD8 on all thymocytes. This will not alter the intensity of a TCR/CD4/MHC Class II signal, and cells receiving this signal will continue development to the CD4 SP lineage. However, down-regulation of CD8 diminishes (interrupts) a TCR/CD8/MHC Class I signal, an experience that sends a cell toward the CD8 lineage. IL-7 signals are required to “seal” CD8 lineage commitment. the strength and duration of the T-cell receptor/coreceptor signal experienced by a thymocyte play a more important role in dictating cell fate than its specificity for MHC Class I or Class II. In fact, thymocytes whose TCR had a known preference of MHC Class II could be coaxed into the CD8⫹ T-cell fate simply by weakening the CD4/Class II interaction (e.g., by mutating kinases downstream of TCR signaling so that they signaled more or less effectively). By inhibiting TCR signaling, investigators could coax cells that normally commit to the CD4 lineage to become CD8⫹ cells. By enhancing TCR signaling, investigators could coax MHC Class I restricted cells to become CD4⫹. These data suggested that stronger positive selecting TCR signals resulted in CD4 lineage commitment, and weaker positive selecting TCR signals resulted in CD8 lineage commitment. They were consistent with the observation that the intracellular tail of CD4 interacts more effectively with the tyrosine kinase lck than the intracellular tail of CD8. Therefore, a TCR/CD4 coengagement is likely to generate stronger signals than TCR/CD8 and result in CD4⫹ T-cell commitment. 316 PA R T I V | Adaptive Immunity: Development The strength of signal model represented an advance in our understanding, but it, too, is likely to be too simplistic. Alfred Singer and colleagues have proposed the kinetic signaling model, which incorporates historical data, our improved understanding of changes in CD8 expression during positive selection, as well as recent advances in the understanding of the complexity of coreceptor gene silencing (Figure 9-9c). Briefly, they propose that thymocytes commit to the CD4⫹ T-cell lineage if they receive a continuous signal in response to TCR/coreceptor engagement, but commit to the CD8 lineage if the TCR signal is interrupted. We now know that all CD4⫹CD8⫹ thymocytes down-regulate surface levels of CD8 in response to positive selection. Given this response, only MHC Class II restricted T cells will maintain continuous TCR/CD4/MHC Class II interaction, and therefore develop to the CD4 lineage. However, with the loss of CD8 expression, MHC Class I restricted T cells will lose the ability to maintain TCR/CD8/MHC Class I interactions. Singer et al. provide evidence that these thymocytes, interrupted from their TCR/ coreceptor engagement, are subsequently rescued by IL-7, which facilitates their commitment to the CD8⫹ lineage. Transcription factors that specifically regulate development to the CD4 and CD8 lineage have recently been identified. At present, ThPok and Runx3 are taking center stage as transcriptional factors required for CD4 and CD8 commitment, respectively, although others also play a role. Both ThPok and Runx3 act, at least in part, by suppressing genes involved in differentiation to the other lineage. ThPok inhibits expression of genes that regulate CD8 differentiation, including Runx3. Runx3 inhibits expression of CD4, itself, as well as ThPok. This is just the beginning of our understanding of the transcriptional networks that regulate CD4 and CD8 developmental programs. Likely to play a role, too, are miRNAs; some have already been implicated in CD8 singlepositive T-cell differentiation. There is clearly more to come. Double-Positive Thymocytes May Commit to Other Types of Lymphocytes Small populations of DP thymocytes can also commit to other T-cell types, including the NK T cell, regulatory T cell, and intraepithelial lymphocyte (IEL) lineages. NKT cells (which include mature cells that express only CD4 and cells that have lost both CD4 and CD8) play a role in innate immunity (see Chapter 5) and express a TCR that includes an invariant TCR␣ chain (V␣14). For this reason, NKT cells are sometimes referred to as iNKT cells for invariant. Their invariant TCR interacts not with classical MHC, but with the related molecule CD1, which presents glycolipids, not peptides (see Chapter 8). Intraepithelial lymphocytes (IEL), most of which are CD8⫹, also have features of innate immune cells and patrol mucosal surfaces. Regulatory T cells (TREG), another CD4⫹ subset discussed in more detail below, quench adaptive immune reactions. All three subpopulations are thought to develop from DP thymocytes in response to autoreactive, high-affinity TCR interactions— the same interactions, in fact, that mediate negative selection. What determines whether a thymocyte undergoes negative selection or an alternative developmental pathway remains a topic of much interest. Exit from the Thymus and Final Maturation Once a thymocyte successfully passes through selection and makes a lineage decision, it enters a quiescent state and leaves the thymus. The cellular and molecular basis for thymocyte egress was unknown until we gained a better understanding of the receptors and ligands that regulate cell migration. As mentioned above, one set of investigators discovered that thymocytes failed to enter the medulla if they were deficient for the CCR7 chemokine receptor. However, these cells still were able to leave the thymus, indicating that other receptors control thymic exit. The identity of this receptor was discovered when investigators found that few if any T cells made it out of the thymus in a sphingosine-1phosphate receptor (SIPR) knockout mouse. Current observations suggest that a cascade of events controls these final stages of maturation: positive selecting TCR signals up-regulate Foxo1, a transcription factor that controls expression of several genes related to T-cell function. Foxo1 regulates expression of Klf2, which, in turn, up-regulates SIPR. Foxo1 also up-regulates both IL-7R, which helps to maintain mature T-cell survival, and CCR7, the chemokine receptor that directs mature T-cell traffic to the lymph nodes. Mature T cells that exit the thymus are referred to as recent thymic emigrants (RTEs). It is now clear that recent thymic emigrants are not as functionally mature as most naïve T cells in the periphery: they do not proliferate or secrete cytokines as vigorously in response to T-cell receptor stimulation. RTEs also can be distinguished from the majority of peripheral naïve T cells because their levels of expression of several surface proteins (e.g., the maturation markers HSA and Qa-2, as well as the IL-7 receptor) are more similar to their immature thymocyte ancestors than their fully mature T-cell descendants. Investigators are particularly interested in RTEs because they are an important source of mature T cells in individuals who are lymphopenic (i.e., have a reduced pool of functional lymphocytes), including people who have undergone chemotherapy, newborns, and the aged. Although research is ongoing, studies suggest that the final maturation of these cells is influenced by their interactions with both MHC and nonMHC ligands in secondary lymphoid organs. Other Mechanisms That Maintain Self-Tolerance As we have seen, negative selection of thymocytes can rid a developing T-cell repertoire of cells that express a high affinity for both ubiquitous self-antigens and, thanks to the | T-Cell Development CHAPTER 9 317 and thwart autoimmune reactions. What determines whether a self-reactive thymocyte dies or differentiates into a TREG cell is still an open question. Investigators are currently trying to understand if the choice is made based on subtle differences in affinity for self or on differences in their maturation state when they receive a high-affinity signal. Recent work suggests that TREGs develop in a unique microenvironmental niche within the thymus, and that the available space for developing cells in this niche is limited. These findings suggest that thymocytes that commit to the regulatory T-cell lineage are likely to receive unique stimulatory signals. These natural TREGs share the periphery with induced TREGs that can develop from conventional mature T cells that are exposed to TGF- and IL-10 cytokines (see Chapter 11). Animal studies show that members of the FoxP3⫹ TREG population inhibit the development of autoimmune diseases such as experimentally induced inflammatory bowel disease, experimental allergic encephalitis, and autoimmune diabetes. Suppression by these regulatory cells is antigen specific because it depends on activation through the T-cell receptor. Exactly how TREGs quench responses is still debated, although they probably do so via a variety of means: directly inhibiting an antigen-presenting cell’s ability to activate T cells, directly killing T cells, indirectly inhibiting T-cell activity by secreting inhibitory cytokines IL-10 and TGF-, and/or depleting the local environment of stimulatory cytokines such as IL-2 (Figure 9-10). activity of AIRE, tissue-specific antigens. However, negative selection in the thymus is not perfect. Autoreactive T cells do escape, either because they have too low an affinity for self to induce clonal deletion, or they happen not to have browsed the “right” tissue-specific antigen/MHC combination. The body has evolved several other mechanisms to avoid autoimmunity, including what has become a major focus of interest for immunologists: the development in both the thymus and the periphery of a fascinating group of cells now known as regulatory T cells. TREG Cells Negatively Regulate Immune Responses Regulatory T cells (TREG cells) can inhibit the proliferation of other T-cell populations in vitro, effectively suppressing autoreactive immune responses. They express on their surface CD4 as well as CD25, the ␣ chain of the IL-2 receptor. However, TREG cells are more definitively identified by their expression of a master transcriptional regulator, FoxP3, the expression of which is necessary and sufficient to induce differentiation to the TREG lineage. TREG cells can develop in the thymus and appear to represent an alternative fate for autoreactive T cells. As we have seen, most thymocytes that express receptors with high affinity for self-antigen die via negative selection. However, a small fraction appear to commit to the regulatory T-cell lineage and leave the thymus to patrol the body 2 T cell Inhibitory cytokines T cell TGF 1 Cytokine deprivation IL–2R FoxP3 3 TCR APC Inhibiting antigen presenting cells MHC 4 Cytotoxicity T cell FIGURE 9-10 How regulatory T cells inactivate traditional T cells. Some possible mechanisms of TREG activity are illustrated in this schematic. These may all contribute to quelling immune responses in vivo. (1) Cytokine deprivation. TREGs express relatively high levels of high-affinity IL-2 receptors and can compete for the cytokines that activated T cells need to survive and proliferate. (2) Cytokine inhibition. TREGs secrete several cytokines, T cell including IL-10 and TGF-, which bind receptors on activated T cells and reduce signaling activity. (3) Inhibition of antigenpresenting cells. TREGs can interact directly with MHC Class II expressing antigen-presenting cells and inhibit their maturation, leaving them less able to activate T cells. (4) Cytotoxicity. TREGs can also display cytotoxic function and kill cells by secreting perforin and granzyme. 318 PA R T I V | Adaptive Immunity: Development The existence of regulatory T cells that specifically suppress immune responses has clinical implications. Depletion or inhibition of TREG cells before immunization may enhance immune responses to conventional vaccines. Elimination of T cells that suppress responses to tumor antigens may also facilitate the development of antitumor immunity. Conversely, increasing the suppressive activity of regulatory T-cell populations could be useful in the treatment of allergic or autoimmune diseases. The ability to increase the activity of regulatory T-cell populations might also be useful in suppressing organ and tissue rejection. Investigators doing future work on this regulatory cell population will seek deeper insights into the mechanisms by which members of FoxP3⫹ T-cell populations regulate immune responses. They will also make determined efforts to discover ways the activities of these populations can be increased to diminish unwanted immune responses and decreased to promote desirable ones. Peripheral Mechanisms of Tolerance also Protect Against Autoreactive Thymocytes The body has evolved several other mechanisms to manage the autoreactive escapee in the periphery. Briefly, many antigens are “hidden” from autoreactive T cells because only a subset of cells (professional APCs) express the right costimulatory molecules needed to initiate the immune response. Autoreactive naïve T cells that can see an MHC/self-peptide combination on a nonprofessional antigen-presenting cell will not receive the correct costimulatory signals and will not divide or differentiate. For example, if a thymocyte specific for a peptide made by a kidney cell escaped from the thymus, it will not be activated unless that peptide were presented on a professional antigenpresenting cell. Kidney cells do not express the costimulatory ligands required for activating a CD4⫹ or a CD8⫹ T cell. In fact, a high-affinity interaction with MHC/peptide combinations on the surface of kidney cell, in the absence of costimulatory ligands, could result in T-cell anergy—another peripheral tolerance mechanism that is described in more detail in Chapter 11. The clinical consequences of failures of central and peripheral tolerance will be discussed in Chapter 16. Apoptosis As we have seen here, and will also see in Chapter 10, apoptosis features prominently during T- and B-lymphocyte development. T-cell development is a particularly expensive process for the host. An estimated 95% to 98% of all thymocytes do not mature—most die by apoptosis within the thymus either because they fail to make a productive TCR gene rearrangement or because they fail to interact with selfMHC. Some (2%–5%) die, also via apoptosis, because they are negatively selected. Not only does apoptosis shape the developing lymphocyte repertoire; it also regulates immune cell homeostasis by returning myeloid, T-, and B-cell populations to their appro- priate levels after bursts of infection-inspired proliferation, and it is the means by which a cell dies after being targeted by cytotoxic T cells or NK cells. Understanding its fundamental biological features is an important part of an overall understanding of immune cell function. Because apoptosis is such a critical part of lymphocyte development, we will take the opportunity in this chapter to briefly describe the apoptotic process and how it is regulated. Apoptosis Allows Cells to Die without Triggering an Inflammatory Response Apoptosis, or programmed cell death, is an energy-dependent process by which a cell brings about its own demise. It is most often contrasted with necrosis, a form of cell death arising from injury or toxicity (Figure 9-11a). In necrosis, injured cells swell and burst, releasing their contents and possibly triggering a damaging inflammatory response. Apoptotic cells, in contrast, dismantle their contents without disrupting their membranes and induce neighboring cells to engulf them before they can release any inflammatory material. Morphologic changes associated with apoptosis include a pronounced decrease in cell volume; modification of the cytoskeleton, which results in membrane blebbing; a condensation of the chromatin; and degradation of the DNA into fragments (Figure 9-11b). Different Stimuli Initiate Apoptosis, but All Activate Caspases Although a lymphocyte can be signaled to die in several ways, all apoptotic signals ultimately activate a class of proteases called caspases. Caspase activation is common to almost all death pathways known in both vertebrates and invertebrates, demonstrating that apoptosis is an ancient process. Many of its molecular participants have been conserved through evolution (Table 9-2). The role of caspases was first revealed by studies of developmentally programmed cell death in the nematode Caenorhabditis elegans, where the death of cells was shown to be totally dependent on the activity of a gene that encoded a cysteine protease with specificity for aspartic acid residues. We now know that mammals have at least 14 cysteine-aspartic proteases, or caspases, and all apoptotic cell deaths require the activity of at least a subset of these molecules. How do caspases trigger cell death? Apoptotic pathways typically involve two classes of caspases, both of which are maintained in an inactive state until apoptosis is initiated. Initiator caspases are activated by a death stimulus and, in turn, activate effector caspases, which execute the death program by (1) cleaving critical targets necessary for cell survival (e.g., cytoskeletal proteins) as well as (2) activating other enzymes that carry out the dismantling of the cell. Ultimately, the catalytic cascade initiated by caspases induces death via an orderly disassembly of intracellular molecules. It also results in the packaging of cell contents into vesicles that are ultimately engulfed. T-Cell Development | CHAPTER 9 319 (a) NECROSIS APOPTOSIS Mild convolution Chromatin compaction and segregation Condensation of cytoplasm Chromatin clumping Swollen organelles Flocculent mitochondria Nuclear fragmentation Blebbing Apoptotic bodies Disintegration Release of intracellular contents Phagocytosis Apoptotic body Phagocytic cell Inflammation (b) Normal Apoptotic TEM FIGURE 9-11 Morphological changes that occur during SEM apoptosis. (a) Comparison of morphologic changes that occur in apoptosis versus necrosis. Apoptosis, which results in the programmed cell death of hematopoietic cells, does not induce a local inflammatory response. In contrast, necrosis, the process that leads to death of injured cells, results in release of the cells’ contents, which may induce a local inflammatory response. (b) Microscopic images of apoptotic cells. Transmission and scanning electron micrographs (TEM and SEM) of normal and apoptotic thymocytes, as indicated. [Part (b) from B. A. Osborne and S. Smith, 1997, Journal of NIH Research 9:35; courtesy of B. A. Osborne, University of Massachusetts at Amherst.] 320 PA R T I V TABLE 9-2 | Adaptive Immunity: Development Proteins involved in apoptosis Protein Location in cell Function Role in apoptosis Death receptors (e.g., Fas, TNFR) Membrane Activates caspase cascade after binding ligand (e.g., FasL, TNF) Promotes Initiator caspase (e.g., caspase-8, caspase-9) Cytosol Protease; cleaves and activates effector caspases Promotes Effector caspase (e.g., caspase-3) Cytosol, nucleus Protease; cleaves and activates enzymes, cleaves and disassembles structural proteins Promotes Cytochrome c Intermembrane space, mitochondria Participates in formation of apoptosome Promotes Apaf-1 Cytosol Participates in formation of apoptosome Promotes Anti-apoptotic Bcl-2 family members (e.g., Bcl-2, Bcl-xL) Mitochondria, ER Regulates cytochrome c release Inhibits Pro-apoptotic Bcl-2 family members (e.g., Bax, Bak) Mitochondria Regulates cytochrome c release; opposes Bcl-2, Bcl-xL Promotes BH3 proteins (e.g., Bim, Bid, PUMA) Cytosol and mitochondria Opposes activity of antiapoptotic Bcl-2 family members at mitochondria Promotes Stimuli that lead to apoptosis can be divided into two categories based on how they activate the effector caspases. Signals such as radiation, stress, and growth factor withdrawal induce caspase activation and apoptosis via the intrinsic pathway that is mediated by mitochondrial molecules, while membrane-bound or soluble ligands that bind to membrane receptors stimulate caspase activation and apoptosis via the extrinsic pathway. Apoptosis of Peripheral T Cells Is Mediated by the Extrinsic (Fas) Pathway Outside of the thymus, most of the TCR-mediated apoptosis of mature T cells is induced via the extrinsic pathway through membrane-associated death receptors, including Fas (CD95). Repeated or persistent stimulation of peripheral T cells via their T-cell receptors results in the co-expression of both Fas and Fas ligand (FasL) on T cells, ultimately leading to Fas/FasL-mediated death. Apoptosis induced by antigen receptor activation is called activation-induced cell death (AICD) and is a major homeostatic mechanism, helping to reduce the pool of activated T cells after antigen is cleared and helping to remove stray autoreactive T cells that are stimulated by self-antigens. The signaling cascade initiated by Fas (and by all death receptors) leads directly to the activation of the initiator caspase, caspase-8. When FasL binds Fas, procaspase-8, an inactive form of caspase-8, is recruited to the intracellular tail of Fas by an adaptor molecule called FADD (Fas-associated protein with death domain) (Figure 9-12a). The aggregation of procaspase-8 results in its cleavage and conversion to active caspase-8, which will then activate effector caspase-3, which in turn initiates the proteolytic cascade that leads to the death of the cell. The importance of Fas and FasL in the removal of mature activated T cells is underscored by abnormalities in lpr/lpr mice, a naturally occurring loss of function mutation in Fas. When T cells become activated in these mice, the Fas/FasL pathway is not operative, and stimulated T cells continue to proliferate. These mice spontaneously develop strikingly large lymph nodes that are filled with excessive numbers of lymphocytes. Ultimately, they develop autoimmune disease, clearly demonstrating the consequences of a failure to delete activated T cells. An additional spontaneous mutant, gld/gld, has the complementary loss of function. These mice lack functional FasL and display abnormalities very similar to those found in the lpr/lpr mice. Recently, humans with defects in Fas have been reported. As expected, these individuals display characteristics of autoimmune disease (see Clinical Focus Box 9-3). T-Cell Development (a) (b) T cell FasL MHC Fas TCR FADD Bcl-2 Caspase-8 (active) Mitochondrion Bid Released cytochrome c Truncated Bid Procaspase-3 (inactive) Caspase-9 Caspase-3 (active) Apoptosis substrates Apoptosome Active apoptotic effectors Apaf-1 Procaspase-9 Apoptosis FIGURE 9-12 Two pathways to apoptosis in T cells. (a) Activated peripheral T cells are induced to express high levels of Fas and FasL and stimulate the extrinsic apoptotic pathway. FasL induces the trimerization of Fas on a neighboring cell. FasL can also engage Fas on the same cell, resulting in a self-induced death signal. Trimerization of Fas leads to the recruitment of FADD, which leads in turn to the cleavage of associated molecules of procaspase-8 to form active caspase-8. Caspase-8 cleaves procaspase-3, producing active caspase-3, which results in the death of the cell. Caspase-8 can also cleave the Bcl-2 family member Bid to a truncated form that can activate the mitochondrial death pathway. (b) Other signals, such as the engagement of the TCR by peptideMHC complexes on an APC during T-cell development, result in the activation of the intrinsic, mitochondrial death pathway. A key feature of this pathway is the release of cytochrome c from the inner mitochondrial membrane into the cytosol, an event that is regulated by Bcl-2 family members. Cytochrome c interacts with Apaf-1 and subsequently with procaspase-9 to form the active apoptosome. The apoptosome initiates the cleavage of procaspase-3, producing active caspase-3, which initiates the execution phase of apoptosis by proteolysis of substances whose cleavage commits the cell to apoptosis. [Adapted in part from S. H. Kaufmann and M. O. Hengartner, 2001, Trends in Cell Biology 11:526.] CHAPTER 9 321 Fas and FasL are members of a family of related receptor/ ligands including tumor necrosis factor (TNF) and its ligand, TNFR (tumor necrosis factor receptor), which can also induce apoptosis via the activation of caspase-8 followed by the activation of effector caspases such as caspase-3 (see Figure 4-14). TCR-Mediated Negative Selection in the Thymus Induces the Intrinsic (MitochondriaMediated) Apoptotic Pathway Bim Procaspase-8 (inactive) | Fas/FasL interactions do not appear to play a central role in negative selection in the thymus. Instead, TCR-mediated negative-selecting signals in the thymus induce a route to apoptosis in which mitochondria play a central role (Figure 9-12b). In mitochondria-dependent (intrinsic) apoptotic pathways, cytochrome c, which normally resides between the inner and outer mitochondrial membranes, leaks into the cytosol. The release of cytochrome c is regulated by several different protein families. In thymocytes, Bim, a Bcl-2 family member, and Nur77, an orphan nuclear receptor, both play a role. In the cytosol, cytochrome c binds to a protein known as Apaf-1 (apoptotic protease-activating factor 1), which then oligomerizes. Binding of oligomeric Apaf-1 to procaspase-9 results in its transformation to active caspase-9, an initiator caspase. The cytochrome c/Apaf-1/caspase-9 complex, called the apoptosome, proteolytically cleaves procaspase-3, generating active caspase-3, which initiates the cascade of reactions that kills the cell. Interestingly, although over 95% of thymocytes die during development, apoptotic thymocytes are very difficult to find in a normal thymus. Investigators showed that if you inhibit the activity of thymic macrophages, apoptotic thymocytes are abundantly evident. This experiment dramatically underscored both the importance and the efficiency of phagocytes in clearing dying cells from the thymus. Cell death induced by Fas/FasL is swift, with rapid activation of the caspase cascade leading to cell death in 2 to 4 hours. On the other hand, TCR-induced negative selection appears to involve a more circuitous pathway, requiring the activation of several processes, including mitochondrial membrane failure, the release of cytochrome c, and the formation of the apoptosome before caspases become involved. Consequently, TCR-mediated negative selection can take as long as 8 to 10 hours. Bcl-2 Family Members Can Inhibit or Induce Apoptosis The Bcl-2 (B-cell lymphoma 2) family of genes, which regulate cytochrome c release, plays a prominent role in immune cell physiology. They include both anti-apoptotic and proapoptotic proteins that can insert into the mitochondrial membrane. Although their precise mechanism of action is still 322 PA R T I V | Adaptive Immunity: Development CLINICAL FOCUS Failure of Apoptosis Causes Defective Lymphocyte Homeostasis The maintenance of appropriate Normal control Patient A 20% CD4–/CD8+ 1% CD4+/CD8+ 24% CD4–/CD8+ 1% CD4+/CD8+ 4% CD4–/CD8– 75% CD4+/CD8– 43% CD4–/CD8– 32% CD4+/CD8– 103 CD8 numbers of various types of lymphocytes is extremely important to an effective immune system. One of the most important elements in this regulation is apoptosis mediated by the Fas/FasL system. The following excerpts from medical histories show what can happen when this key regulatory mechanism fails. Patient A: A woman, now 43, has had a long history of immunologic imbalances and other medical problems. By age 2, she was diagnosed with CanaleSmith syndrome (CSS), a severe enlargement of such lymphoid tissues as lymph nodes (lymphadenopathy) and spleen (splenomegaly). Biopsy of lymph nodes showed that, in common with many other CSS patients, she had greatly increased numbers of lymphocytes. She had reduced numbers of platelets (thrombocytopenia) and, because her red blood cells were being lysed, she was anemic (hemolytic anemia). The reduction in numbers of platelets and the lysis of red blood cells could be traced to the action of circulating antibodies that reacted with these host components. At age 21, she was diagnosed with grossly enlarged pelvic lymph nodes that had to be removed. Ten years later, she was again found to have an enlarged abdominal mass, which on surgical removal turned out to be a half-pound lymph node aggregate. She has continued to have mild lymphadenopathy and, typical of these patients, the lymphocyte populations of enlarged nodes had elevated numbers of T cells (87% as opposed to a 104 102 101 100 100 101 102 CD4 103 104 100 101 102 CD4 103 104 FIGURE 1 Flow-cytometric analysis of T cells in the blood of CSS patient A and a control subject. Mature T cells are either CD4⫹ or CD8⫹. Although almost all of the T cells in the control subject are CD4⫹ or CD8⫹, the CSS patient shows high numbers of double-negative T cells (43%), which express neither CD4 nor CD8. The percentage of each category of T cells is indicated in the quadrants. [Adapted from M. D. Drappa et al., 1996, New England Journal of Medicine 335:1643.] normal range of 48%–67% T cells). Examination of these cells by flow cytometry and fluorescent antibody staining revealed an excess of double-negative T cells (Figure 1). Also, like many patients with CSS, she has had cancer: breast cancer at age 22 and skin cancer at ages 22 and 41. Patient B: A man who was eventually diagnosed with CSS had severe lymphadenopathy and splenomegaly as an infant and child. He was treated from age 4 to age 12 with corticosteroids and the immunosuppressive drug mercaptopurine. These appeared to help, and the swelling of lymphoid tissues became milder during adolescence and adulthood. At age 42, he died of liver cancer. controversial, the balance of pro-apoptotic and anti-apoptotic family members influences a cell’s response to stress. The Bcl-2 family also includes an important, third group of proteins, the BH3 family, all of which do not insert in the mitochondrial membrane, but instead inhibit the anti-apoptotic Bcl-2 family members, and so are ultimately pro-apoptotic. Intrinsic apoptotic stimuli act, in part, by activating BH3 family members, which are often sequestered from the mito- Patient C: An 8-year-old boy, the son of patient B, was also afflicted with CSS and showed elevated T-cell counts and severe lymphadenopathy at the age of 7 months. At age 2 his spleen became so enlarged that it had to be removed. He also developed hemolytic anemia and thrombocytopenia. However, although he continued to have elevated T-cell counts, the severity of his hemolytic anemia and thrombocytopenia have so far been controlled by treatment with methotrexate, a DNA-synthesis-inhibiting drug used for immunosuppression and cancer chemotherapy. Recognition of the serious consequences of a failure to regulate the number of lymphocytes, as exemplified by chondria in a healthy cell. For example, TCR-mediated negative selection activates Bim, a BH3 family protein that facilitates cytochrome c release (see Figure 9-12b). However, even the extrinsic pathway can activate BH3 family members. For instance, caspase-8 generated by the Fas pathway cleaves the BH3 family member Bid, releasing it from sequestration and allowing it to activate the mitochondrial death pathway. | T-Cell Development CHAPTER 9 323 BOX 9-3 60 Percentage of T cells killed these case histories, emerged from detailed study of several children whose enlarged lymphoid tissues attracted medical attention. In each of these cases of CSS, examination revealed grossly enlarged lymph nodes that were 1 cm to 2 cm in girth and sometimes large enough to distort the local anatomy. In four of a group of five children who were studied intensively, the spleens were so massive that they had to be removed. Even though the clinical picture in CSS can vary from person to person, with some individuals suffering severe chronic affliction and others only sporadic episodes of illness, there is a common feature: a failure of activated lymphocytes to undergo Fas-mediated apoptosis. Isolation and sequencing of Fas genes from a number of patients and more than a hundred controls reveals that CSS patients are heterozygous (fas⫹/⫺) at the fas locus and thus carry one copy of a defective fas gene. A comparison of Fas-mediated cell death in T cells from normal controls who do not carry mutant Fas genes with death induced in T cells from CSS patients shows a marked defect in Fas-induced death (Figure 2). Characterization of the Fas genes so far seen in CSS patients reveals that they have mutations in or around the region encoding the deathinducing domain (the “death domain”) of this protein. Such mutations result in the production of Fas protein that lacks biological activity but still competes with normal Fas molecules for interactions with essential components of the Fas-mediated death pathway. Other mutations have been found in the extracellular domain of Fas, often associated with milder forms of CSS or no disease at all. Normal controls Patient A 40 Patient B 20 16 80 Anti-Fas antibody (ng/ml) 400 FIGURE 2 Fas-mediated killing takes place when Fas is cross-linked by FasL, its normal ligand, or by treatment with anti-Fas antibody, which artificially cross-links Fas molecules. This experiment shows the percentage of dead T cells after induction of apoptosis in T cells from patients and controls by cross-linking Fas with increasing amounts of an anti-Fas monoclonal antibody. T cells from the Canale-Smith patients (A and B) are resistant to Fas-mediated death. [Adapted from M. D. Drappa et al., 1996, New England Journal of Medicine 335:1643.] A number of research groups have conducted detailed clinical studies of CSS patients, and the following general observations have been made: • The cell populations of the blood and lymphoid tissues of CSS patients show dramatic elevations (5-fold to as much as 20-fold) in the numbers of lymphocytes of all sorts, including T cells, B cells, and NK cells. • Most of the patients have elevated levels of one or more classes of immunoglobulin (hyper-gammaglobulinemia). • Immune hyperactivity is responsible for such autoimmune phenomena as the production of autoantibodies against red blood cells, resulting in hemolytic anemia, and a depression in platelet counts due to the activity of antiplatelet autoantibodies. The signature member of this gene family, bcl-2, was discovered near the breakpoint of a chromosomal translocation in a cancer known as human B-cell lymphoma. The translocation moved the bcl-2 gene into the immunoglobulin heavy-chain locus, resulting in transcription of bcl-2 along with the immunoglobulin gene, with consequent overproduction of the encoded Bcl-2 protein by the lymphoma cells. The resulting high levels of Bcl-2 resulted in the accumula- These observations establish the importance of the death-mediated regulation of lymphocyte populations in lymphocyte homeostasis. Such cell death is necessary because the immune response to antigen results in a sudden and dramatic increase in the populations of responding clones of lymphocytes and temporarily distorts the representation of these clones in the repertoire. In the absence of cell death, the periodic stimulation of lymphocytes that occurs in the normal course of life would result in progressively increasing, and ultimately unsustainable, lymphocyte levels. As CSS demonstrates, without the essential culling of lymphocytes by apoptosis, severe and life-threatening disease can result. tion of long-lived B cells and contributed to their transformation into cancerous lymphoma cells. Bcl-2 levels play an important role in regulating the normal life span of various hematopoietic cell lineages, including lymphocytes. A normal adult has about 5 liters of blood, with about 2000 lymphocytes/mm3, for a total of about 1011 circulating lymphocytes. During acute infection, the lymphocyte count increases fourfold or more. The 324 PA R T I V | Adaptive Immunity: Development immune system cannot sustain such a massive increase in cell numbers for an extended period, so the system must eliminate unneeded activated lymphocytes once the antigenic threat has passed. Activated lymphocytes have been found to express lower levels of Bcl-2, and therefore are more susceptible to the induction of apoptotic death than are naïve lymphocytes or memory cells. However, if the lymphocytes continue to be activated by antigen, signals received during activation block the apoptotic signal. As antigen levels subside, the levels of the signals that block apoptosis diminish, and the lymphocytes begin to die by apoptosis. S U M M A R Y ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ Uncommitted white blood cell progenitors enter the thymus from the bone marrow. Notch/Notch ligand interactions are required for T-cell commitment. Immature T cells are called thymocytes, and stages of development can be defined broadly by the expression of the coreceptors CD4 and CD8. The most immature thymocytes express neither coreceptor and are referred to as CD4⫺CD8⫺ or double-negative (DN) cells. These progress to the CD4⫹CD8⫹ (double positive or DP) stage, which in turn mature to the CD4⫹CD8⫺ or CD4⫺CD8⫹ (single positive or SP) stages. DN thymocytes progress through four stages of development (DN1-DN4) defined by CD44, CD25, and c-Kit expression. During these stages they proliferate and rearrange the TCR, ␦, and ␥ antigen receptor genes. Thymocytes that rearrange ␦ and ␥ receptor genes successfully mature to the TCR␥␦ lineage. Those thymocytes that successfully rearrange the ␤ receptor chain are detected by a process called ␤-selection. ␤-selection is initiated by assembly of the TCR␤ protein with a surrogate, invariant TCR␣ chain and CD3 complex proteins. Assembly of this pre-TCR results in commitment to the TCR␣␤ lineage, another burst of proliferation, maturation to the DP stage of development, and initiation of TCR␣ rearrangement. DP thymocytes are the most abundant subpopulation in the thymus and are the first cells to express a mature TCR␣␤ receptor. They are the targets of positive and negative selection, which are responsible for self-tolerance and self-MHC restriction, respectively. Positive and negative selection are regulated by the affinity of a DP thymocyte’s TCR for MHC/self-peptide complexes expressed by the thymic epithelium. High-affinity TCR/MHC/peptide interactions result in negative selection, typically by initiating apoptosis (clonal deletion). Low/intermediate-affinity signals result in positive selection and initiate a maturation program to the helper CD4⫹ or cytotoxic CD8⫹ single-positive lineages. The large majority of thymocytes (95%) do not interact with any MHC/self-peptides expressed by the thymic epithelium and die by neglect. ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ The distinction among TCR signaling cascades that activate positive versus negative selection is not fully understood but may involve differences in MAPK and Ca2⫹ signaling. Positively selected DP thymocytes have to decide whether to become helper CD4⫹ or cytotoxic CD8⫹ T cells. This process, called lineage commitment, appears to depend on continuity of the signals positively selected DP thymocytes receive through their TCR. Thymocytes whose TCR preferentially interacts with MHC Class II generate a continuous signal that initiates a CD4⫹ helper T-cell developmental program. Thymocytes whose TCR preferentially interacts with MHC Class I cannot generate a continuous signal because CD8 surface levels are down-regulated in response to positive selection. The interruption in signaling, followed by further stimulation, initiates a CD8 developmental program. Transcription factors Th-Pok and Runx3 play important roles in CD4⫹ helper cell and CD8⫹ cytotoxic cell commitment. They work in part by reciprocally regulating each other. Positively selected thymocytes migrate from the thymic cortex to the thymic medulla via interactions with the CCR7 chemokine receptor. These cells, in transition from the DP to the SP stage of development, are negatively selected for tissue-specific antigens in the medulla. This is the second opportunity to remove autoreactive T cells from the developing repertoire. The thymic cortex and thymic medulla carry out distinct functions in the thymus. Medullary epithelial cells, but not cortical epithelial cells, express the transcription factor AIRE, which is responsible for their unique capacity to express tissue-specific antigens. Fully mature thymocytes exit the thymus via interactions via the sphingosine-1-phosphate receptor (S1PR) and undergo final functional maturation in peripheral lymphoid tissues. A small percentage of cells also develop within the thymus to other cell lineages, including NK T cells, IELs, and regulatory T cells. Regulatory T cells that develop in the thymus are called natural TREGs. They share the periphery with regulatory T cells T-Cell Development ■ ■ that develop from conventional T cells (induced TREGs) and play an important role in inhibiting autoimmune responses. Apoptosis, a process of cell death that is internally initiated and highly regulated, has a major influence on the shape of the T-cell repertoire. Thymocytes that do not pass -selection (and do not successfully express a TCR␥␦ complex) die by apoptosis; 95% of developing DP thymocytes also die by apoptosis, the vast majority because they do not receive sufficient TCR signaling (death by neglect) and a smaller population because they receive too high a level of TCR signaling (negative selection). The process by which a cell undergoes apoptosis signaling is evolutionarily conserved and inspired by many different receptor interactions in most if not all cell types in the body. ■ ■ ■ | CHAPTER 9 325 Apoptotic stimuli either activate death receptors (extrinsic pathway) or mitochondrial cytochrome c release (intrinsic pathway). Both pathways ultimately activate effector caspases, which finalize cell death events. Apoptotic cells are rapidly engulfed by neighboring phagocytes and, for this reason, are difficult to detect in living tissues. Bcl-2 family members include both anti-apoptotic and pro-apoptotic family members that reside in the mitochondrial membrane and regulate cytochrome c release. A third set of family members, the pro-apoptotic BH3 proteins, can be activated by apoptotic stimuli, favoring a shift in balance of Bcl-2 family member activity at the mitochondrial membrane toward cytochrome c release. R E F E R E N C E S Alam, S., et al. 1999. Qualitative and quantitative differences in T cell receptor binding of agonist and antagonist ligands. Immunity 10:227–237. Germain, R. 2008. Special regulatory T-cell review: A rose by any other name: from suppressor T cells to TREGs, approbation to unbridled enthusiasm. Immunology 123:20–27. Anderson, M., et al. 2002. Projection of an immunological self shadow within the thymus by the aire protein. Science 298:1395–1401. He, X., et al. 2005. The zinc finger transcription factor Th-POK regulates CD4 versus CD8 T-cell lineage commitment. Nature 433:826–833. Baldwin, T., K. Hogquist, and S. Jameson. 2004. The fourth way? Harnessing aggressive tendencies in the thymus. Journal of Immunology 173:6515–6520. Hogquist, K., T. Baldwin, and S. Jameson. 2005. Central tolerance: Learning self-control in the thymus. Nature Reviews. Immunology 5:772–782. Baldwin, T., M. Sandau, S. Jameson, and K. Hogquist. 2005. The timing of TCR alpha expression critically influences T cell development and selection. Journal of Experimental Medicine 202:111–121. Hogquist, K., M. Gavin, and M. Bevan. 1993. Positive selection of CD8⫹ T cells induced by major histocompatibility complex binding peptides in fetal thymic organ culture. Journal of Experimental Medicine 177:1469–1473. Bouillet, P., et al. 2002. BH3-only Bcl-2 family member Bim is required for apoptosis of autoreactive thymocytes. Nature 415:922–926. Hogquist, K., S. Jameson, and M. Bevan. 1995. Strong agonist ligands for the T cell receptor do not mediate positive selection of functional CD8⫹ T cells. Immunity 3:79-86. Carpenter, A., and R. Bosselut. 2010. Decision checkpoints in the thymus. Nature Immunology 11:666–673. Hogquist, K., et al. 1994. T cell receptor antagonist peptides induce positive selection. Cell 76:17–27. Caton, A., et al. 2004. CD4(⫹) CD25(⫹) regulatory T cell selection. Annals of the New York Academy of Sciences 1029:101–114. Kappes, D., and X. He. 2006. ole of the transcription factor ThPOK in CD4:CD8 lineage commitment. Immunological Reviews 209:237–252. Ceredig, R., and T. Rolink. 2002. A positive look at doublenegative thymocytes. Nature Reviews Immunology 2:888–897. Drennan, M., D. Elewaut, and K. Hogquist. 2009. Thymic emigration: sphingosine-1-phosphate receptor-1-dependent models and beyond. European Journal of Immunology 39:925–930. Kisielow, P., H. Blüthmann, U. Staerz, M. Steinmetz, and H. von Boehmer, H. 1988. Tolerance in T-cell-receptor transgenic mice involves deletion of nonmature CD4⫹8⫹ thymocytes. Nature 333:742–746. Gascoigne, N. 2010. CD8⫹ thymocyte differentiation: T cell two-step. Nature Immunology 11:189–190. Klein, L., M. Hinterberger, G. Wirnsberger, and B. Kyewski. 2009. Antigen presentation in the thymus for positive selection and central tolerance induction. Nature Reviews. Immunology 9:833–844. Germain, R. 2002. T-cell development and the CD4-CD8 lineage decision. Nature Reviews. Immunology 2:309–322. Kyewski, B., and P. Peterson. 2010. Aire: Master of many trades. Cell 140:24–26. 326 PA R T I V | Adaptive Immunity: Development Li, R., and D. Page. 2001. Requirement for a complex array of costimulators in the negative selection of autoreactive thymocytes in vivo. Journal of Immunology 166:6050–6056. Uematsu, Y., et al. 1988. In transgenic mice the introduced functional T cell receptor beta gene prevents expression of endogenous beta genes. Cell 52:831–841. Marrack, P., L. Ignatowicz, J. Kappler, J. Boymel, and J. Freed. 1993. Comparison of peptides bound to spleen and thymus class II. Journal of Experimental Medicine 178:2173–2183. Venanzi, E., C. Benoist, and D. Mathis. 2004. Good riddance: Thymocyte clonal deletion prevents autoimmunity. Current Opinion in Immunology 16:197–202. Marrack, P., J. McCormack, and J. Kappler. 1989. Presentation of antigen, foreign major histocompatibility complex proteins and self by thymus cortical epithelium. Nature 338:503–505. von Boehmer, H., H. Teh, and P. Kisielow. 1989. The thymus selects the useful, neglects the useless and destroys the harmful. Immunology Today 10:57–61. Mathis, D., and C. Benoist. 2009. Aire. Annual Review of Immunology 27:287–312. Wang, L., and R. Bosselut. 2009. CD4-CD8 lineage differentiation: Thpok-ing into the nucleus. Journal of Immunology 183:2903–2910. McNeil, L., T. Starr, and K. Hogquist. 2005. A requirement for sustained ERK signaling during thymocyte positive selection in vivo. Proceedings of the National Academy of Sciences of the United States of America 102:13,574–13,579. Mohan, J., et al. 2010. Unique autoreactive T cells recognize insulin peptides generated within the islets of Langerhans in autoimmune diabetes. Nature Immunology 11:350–354. Page, D., L. Kane, J. Allison, and S. Hedrick. 1993. Two signals are required for negative selection of CD4⫹CD8⫹ thymocytes. Journal of Immunology 151:1868–1880. Park, J., et al. 2010. Signaling by intrathymic cytokines, not T cell antigen receptors, specifies CD8 lineage choice and promotes the differentiation of cytotoxic-lineage T cells. Nature Immunology 11:257–264. Pui, J., et al. 1999. Notch1 expression in early lymphopoiesis influences B versus T lineage determination. Immunity 11:299–308. Punt, J., B. Osborne, Y. Takahama, S. Sharrow, and A. Singer. 1994. Negative selection of CD4⫹CD8⫹ thymocytes by T cell receptor-induced apoptosis requires a costimulatory signal that can be provided by CD28. Journal of Experimental Medicine 179:709–713. Wang, L., et al. 2008. Distinct functions for the transcription factors GATA-3 and ThPOK during intrathymic differentiation of CD4(⫹) T cells. Nature Immunology 9:1122–1130. Wilson, A., H. MacDonald, and F. Radtke. 2001. Notch 1-deficient common lymphoid precursors adopt a B cell fate in the thymus. Journal of Experimental Medicine 194:1003–1012. Zúñiga-Pflücker, J. 2004. T-cell development made simple. Nature Reviews. Immunology 4:67–72. Useful Web Sites http://www.bio.davidson.edu/courses/immunology/ Flash/Main.htm A Dr. Victor Lemas–generated animation of positive and negative selection events in the thymus, used at Davidson College. http://www.bio.davidson.edu/courses/movies. html A full list of animations assembled by and in many Sambandam, A., et al. 2005. Notch signaling controls the generation and differentiation of early T lineage progenitors. Nature Immunology 6:663–670. cases generated by individuals associated with Davidson College. Schmitt, T., and J. Zúñiga-Pflücker. 2002. Induction of T cell development from hematopoietic progenitor cells by deltalike-1 in vitro. Immunity 17:749–756. http://bmc.med.utoronto.ca/student_video_ gallery.php?titleimages/stories/videos/Janice_ Yau&h360&w640&m4v1 Another animation of Singer, A. 2010. Molecular and cellular basis of T cell lineage commitment: An overview. Seminars in Immunology 22:253. T-cell development generated as part of a master’s project by Janice Yau. This is from an impressive site that features multiple video projects generated by master’s students in the University of Toronto at Mississauga’s Biomedical Communications Graduate Program. Stadinski, B., et al. 2010. Diabetogenic T cells recognize insulin bound to IAg7 in an unexpected, weakly binding register. Proceedings of the National Academy of Sciences of the United States of America 107:10,978–10,983. Starr, T., S. Jameson, and K. Hogquist. 2003. Positive and negative selection of T cells. Annual Review of Immunology 21:139–176. http://bio-alive.com/categories/apoptosis.htm Videos of cells undergoing apoptosis. www.celldeath.de/encyclo/aporev/aporev.htm Sun, G., et al. 2005. The zinc finger protein cKrox directs CD4 lineage differentiation during intrathymic T cell positive selection. Nature Immunology 6:373–381. Detailed online review of apoptosis biology. Teh, H., et al. 1988. Thymic major histocompatibility complex antigens and the alpha beta T-cell receptor determine the CD4/CD8 phenotype of T cells. Nature 335:229–233. Access to original Kerr, Wyllie, and Currie British Journal of Cancer 1972 article where the term apoptosis was coined. www.ncbi.nlm.nih.gov/pmc/articles/PMC2008650 T-Cell Development CHAPTER 9 327 Q U E S T I O N S 2. Susceptibility to many autoimmune diseases has been linked to MHC gene variants. One of the best examples of such a linkage is provided by multiple sclerosis (MS), a human autoimmune disease caused by autoreactive T cells whose activity ultimately damages the myelin sheaths around neurons. Susceptibility to MS has been consistently associated with variants in the HLA-DR2 gene. Although this link was first recognized in 1972, we still don’t fully understand the basis for this susceptibility. One perspective on the reasons for the link between MHC variations and autoimmune disease was offered in a recent review article. The authors state, “The mechanisms underlying MHC association in autoimmune disease are not clearly understood. One long-held view suggests a breakdown in immunological tolerance to self-antigens through aberrant class II presentation of self or foreign peptides to autoreactive T lymphocytes. Thus, it seems likely that specific MHC class II alleles determine the targeting of particular autoantigens resulting in disease-specific associations.” (Fernando, M. M. A., et al. 2008 Defining the role of the MHC in autoimmunity: A review and pooled analysis. PLoS Genet 4:4: e1000024. doi:10.1371/journal.pgen.1000024.) a. Paraphrase this perspective using your own words. What, specifically, might the authors mean by “aberrant class II presentation . . . to autoreactive T lymphocytes”? b. (Advanced question.) Although this speculation has some merit, it does not resolve all questions. Why? Pose one question that this explanation inspires or does not answer. c. (Very advanced question.) Offer one addition to the explanation (or an alternative) that helps resolve the question you posed above. 3. Over a period of several years, a group of children and ado- lescents are regularly dosed with compound X, a lifesaving drug. However, in addition to its beneficial effects, it was found that this drug interferes with Fas-mediated signaling. a. What clinical manifestations of this side effect of compound X might be seen in these patients? b. If white blood cells from an affected patient are stained with a fluorescein-labeled anti-CD4 and a phycoerythrin-labeled anti-CD8 antibody, what might be seen in the flow-cytometric analysis of these cells? What pattern would be expected if the same procedure were performed on white blood cells from a healthy control? STUDY QUESTIONS 1. Each of the following statements is false. Correct them (and explain your correction[s]). Knockout mice lacking class I MHC molecules fail to produce CD4⫹ mature thymocytes. -selection initiates negative selection. 2. Whereas the majority of T cells in our bodies express an ␣␤ TCR, up to 5% of T cells express the ␥␦ TCR instead. Explain the difference in antigen recognition between these two cell types. 3. You have fluorescein-labeled anti-CD4 and phycoerythrin- labeled anti-CD8. You use these antibodies to stain thymocytes and lymph-node cells from normal mice and from RAG-1 knockout mice. In the forms below, draw the FACS plots that you would expect. Thymus Normal mice RAG-1 knockout mice CD4 basis and has been linked to many different gene loci. Identify three possible genes other than Fas and AIRE (whose connection to autoimmunity have been explicitly described in this chapter) that could be involved in an increased susceptibility to autoimmune disease. Explain your reasoning. CD8 Normal mice CD8 Lymph node RAG-1 knockout mice CD4 1. The susceptibility to autoimmune diseases often has a genetic Negative selection to tissue-specific antigens occurs in the cortex of the thymus. Most thymocytes die in the thymus because they are autoreactive. The extrinsic pathway of apoptosis never activates the intrinsic pathway. Cytochrome c is an important downstream molecule in the extrinsic apoptotic pathway. Bcl-2 enhances the activity of Bax and therefore inhibits apoptosis. CD4 CLINICAL FOCUS QUESTIONS CD4 S T U D Y | CD8 CD8 4. What stages of T-cell development (DN1, DN2, DN3, DN4, DP, CD4 SP, or CD8 SP) would be affected in mice with the following genetic modifications? Justify your answers. a. Mice that do not express MHC Class II. b. Mice that do not express AIRE. c. Mice that do not express the TCR␣ chain. 5. You stain thymocytes with PE conjugated anti-CD3 and FITC conjugated anti-TCR␤. Most cells stain with both. However, you find a proportion of cells that stain with neither antibody. You also find a small population that stain with anti-CD3, but not with anti-TCR␤. What thymocyte populations might each of these populations represent? 328 PA R T I V | Adaptive Immunity: Development k 6. You immunize an H-2 mouse with chicken ovalbumin (a protein against which the mouse will generate an immune response) and isolate a CD4⫹ mature T cell specific for an ovalbumin peptide. You clone the ␣␤ TCR genes from this cell line and use them to prepare transgenic mice with the H-2k or H-2d haplotype. a. What approach can you use to distinguish immature thymocytes from mature CD4⫹ thymocytes in the transgenic mice? b. Would thymocytes from a TCR transgenic mouse of the H-2k background have a proportion of CD4⫹ thymocytes that is higher, lower, or the same as a wild-type mouse? c. Would thymocytes from a TCR transgenic mouse of the H-2d background have a proportion of CD4⫹ thymocytes that is higher, lower, or the same as a wild-type mouse? Speculate and explain your answer. d. You find a way to “make” the medullary epithelium of an H-2k TCR transgenic mouse express and present the ovalbumin peptide for which your T cell is specific. What would the CD4 versus CD8 profile of a TCR transgenic thymus look like in these mice? You also find a way to “make” the cortical epithelium express this ovalbumin peptide in its MHC Class II dimer. What might the CD4 versus CD8 profile of this TCR transgenic thymus look like? 7. In his classic chimeric mouse experiments, Zinkernagel took bone marrow from a mouse of one MHC haplotype (mouse 1) and a thymus from a mouse of another MHC haplotype (mouse 2) and transplanted them into a third mouse, which was thymectomized and lethally irradiated. He then immunized this reconstituted mouse with the lymphocytic choriomeningitis virus (LCMV) and examined the activity of the mature T cells isolated from the spleen and lymph nodes of the mouse. He was specifically interested to see if the mature CD8⫹ T cells in these mice could kill target cells infected with LCMV with the MHC haplotype of mouse 1, 2, or 3. The results of two such experiments using H-2b strain C57BL/6 mice and H-2d strain BALB/c mice as bone marrow and thymus donors, respectively, are shown in the following table. Lysis of LCMVinfected target cells Bone Thymectomized marrow Thymus x-irradiated Experiment donor donor recipient H-2d H-2k H-2b A C57BL/6 BALB/ c (H-2d) (H-2b) C57BL/6 ⫻ BALB/c ⫹ ⫺ ⫺ B BALB/c (H-2d) C57BL/6 C57BL/6 ⫻ (H-2b) BALB/c ⫺ ⫺ ⫹ b a. Why were the H-2 target cells not lysed in experiment A but lysed in experiment B? k b. Why were the H-2 target cells not lysed in either experiment? ⫹ k 8. You have a CD8 CTL clone (from an H-2 mouse) that has a T-cell receptor specific for the H-Y antigen. You clone the ␣␤ TCR genes from this cloned cell line and use them to prepare transgenic mice with the H-2k or H-2d haplotype. a. How can you distinguish the immature thymocytes from the mature CD8⫹ thymocytes in the transgenic mice? b. For each of the following transgenic mice, indicate with (1) or (2) whether the mouse would have immature double-positive and mature CD8⫹ thymocytes bearing the transgenic T-cell receptor: H-2k female, H-2k male, H-2d female, H-2d male. k c. Explain your answers for the H-2 transgenics. d d. Explain your answers for the H-2 transgenics. 9. To demonstrate positive thymic selection experimentally, researchers analyzed the thymocytes from normal H-2b mice, which have a deletion of the class II IE gene, and from H-2b mice in which the class II IA gene had been knocked out. a. What MHC molecules would you find on antigenpresenting cells from the normal H-2b mice? b. What MHC molecules would you find on antigenpresenting cells from the IA knockout H-2b mice? ⫹ ⫹ c. Would you expect to find CD4 T cells, CD8 T cells, or both in each type of mouse? Why? 10. You wish to determine the percentage of various types of thymocytes in a sample of cells from mouse thymus using the indirect immunofluorescence method. a. You first stain the sample with goat anti-CD3 (primary antibody) and then with rabbit FITC-labeled antigoat Ig (secondary antibody), which emits a green color. Analysis of the stained sample by flow cytometry indicates that 70% of the cells are stained. Based on this result, how many of the thymus cells in your sample are expressing antigenbinding receptors on their surface? Would all be expressing the same type of receptor? Explain your answer. What are the remaining unstained cells likely to be? ⫹ b. You then separate the CD3 cells with the fluorescenceactivated cell sorter (FACS) and restain them. In this case, the primary antibody is hamster anti-CD4, and the secondary antibody is rabbit PE-labeled antihamster Ig, which emits a red color. Analysis of the stained CD3⫹ cells shows that 80% of them are stained. From this result, can you determine how many TC cells are present in this sample? If yes, then how many TC cells are there? If no, what additional experiment would you perform to determine the number of TC cells that are present? 10 B-Cell Development M illions of B lymphocytes are generated in the bone marrow every day and exported to the periphery. The rapid and unceasing generation of new B cells occurs in a carefully regulated sequence of events. Cell transfer experiments, in which genetically marked donor hematopoietic stem cells (HSCs) are injected into an unmarked recipient, have indicated that B-cell development from HSC to mature B cell takes from 1 to 2 weeks; donor-derived mature B cells can be detected in the recipient by 2 weeks following transfer of HSCs into recipient mice. B-cell development begins in the bone marrow with the asymmetric division of an HSC and continues through a series of progressively more differentiated progenitor stages to the production of common lymphoid progenitors (CLPs), which can give rise to either B cells or T cells (see Overview Figure 10-1). Progenitor cells destined to become T cells migrate to the thymus where they complete their maturation (see Chapter 9); the majority of those that remain in the bone marrow become B cells. As differentiation proceeds, the developing B cell expresses on its cell surface a precisely calibrated sequence of cell-surface receptor and adhesion molecules. Some of the signals received from these receptors induce the differentiation of the developing B cell; others trigger its proliferation at particular stages of development and yet others direct its movements within the bone marrow environment. These signals collectively allow differentiation of the CLP through the early B-cell stages to form the immature B cell that leaves the marrow to complete its differentiation in the spleen. For the investigator, the expression of different cellsurface molecules at each stage of B-cell maturation provides an invaluable experimental tool with which to recognize and isolate B cells poised at discrete points in their development. The primary function of mature B cells is to secrete antibodies that protect the host against pathogens, and so one major focus of those studying B-cell differentiation is the analysis of the timing and order of rearrangement and expression of immunoglobulin receptor heavy- and light-chain genes. Recall from Chapter 7 that immunoglobulin gene rearrangements begin with B cells at different stages of development seek contact with stromal cells expressing CXCL12 (pre-pro-B cells, left) or IL-7 (pro-B cells, right). [Tokoyoda et al. 2004. Cellular Niches Controlling B Lymphocyte Behavior within Bone Marrow during Development. Immunity Vol. 20, Issue 6, 707–718. © 2004 Elsevier Ltd. All rights reserved.] ■ The Site of Hematopoiesis ■ B-Cell Development in the Bone Marrow ■ ■ The Development of B-1 and Marginal-Zone B Cells Comparison of B- and T-Cell Development heavy-chain D to JH gene-segment rearrangement, followed by the stitching together of the  heavy-chain VH and DJH segments. These rearrangements culminate in the cell-surface expression of the pre-B-cell receptor during the pre-B-cell stage, in which the rearranged heavy chain is expressed in combination with the surrogate light chain. Rearrangement of the light chain is initiated after several rounds of division of cells bearing the pre-BCR. Like T cells, developing B cells must solve the problem of creating a repertoire of receptors capable of recognizing an extensive array of antigens, while ensuring that self-reactive B cells are either eliminated by apoptosis or rendered functionally unreactive or anergic. However, unlike T-cell receptors, B-cell receptors are not constrained by the need to be MHC restricted. Again, unlike T-cell maturation, B-cell development is almost complete by the time the B cell leaves the bone marrow; in mammalian systems there is no thymic equivalent in which B-cell development is accomplished. Instead, immature B cells are released to 329 330 PA R T I V | Adaptive Immunity: Development 10-1 OVERVIEW FIGURE B-Cell Development Begins in the Bone Marrow and Is Completed in the Periphery Early ProB ProB or preB • DJ H chain • Complete recombination VDJ H chain • Start of VDJ H chain recombination recombination • Clonal expansion • VJ L chain recombination Immature B Negative selection • Deletion • Receptor editing Endogenous antigen Bone endosteum Central sinus Bone marrow Transitional-2 Mature B IgM IgD Transitional-1 Spleen B-cell development begins with a hematopoietic stem cell (HSC) and passes through progressively more delimited progenitor-cell stages until it reaches the pro-B cell stage. At this stage, the precursor cell is irreversibly committed to the B-cell lineage and the the periphery, where they complete their developmental program in the spleen. In this chapter, we will follow B-cell development from its earliest stages in the primary lymphoid organs to the generation of fully mature B cells in the secondary lymphoid tissues. As for T cells, multiple B-cell subsets exist, and we will briefly address how the process of differentiation of the minority B-1 and marginal-zone (MZ) B-cell subsets differs from the developmental program followed by the predominant B-2 B-cell subset. We will conclude with a brief comparison of the maturational processes of T and B lymphocytes. The Site of Hematopoiesis In adult animals, hematopoiesis, the generation of blood cells, occurs in the bone marrow; the HSCs in the marrow are the source of all blood cells of the erythroid, myeloid, and recombination of the immunoglobulin genes begins. Once the completed immunoglobulin is expressed on the cell surface, the immature B cell, now a transitional B cell, leaves the bone marrow to complete its maturation in the spleen. lymphoid lineages (Chapter 2). Various non-hematopoietic cells in the bone marrow express cell-surface molecules and secrete hormones that guide hematopoietic cell development. Developing lymphocytes move within the bone marrow as they mature, thus interacting with different populations of cells and signals at various developmental stages. However, fetal animals face particular challenges to their developing immune systems; how can they generate blood cells when their bones are still not yet fully developed? The Site of B-Cell Generation Changes during Gestation Hematopoiesis is a complex process in the adult animal, and during fetal maturation additional challenges must be met. Red blood cells must be quickly generated de novo in order to provide the embryo with sufficient oxygen, and HSCs must proliferate at a rate sufficient to populate the adult as well as provide for the hematopoietic needs of the maturing fetus. Furthermore, since the bone marrow appears relatively late in B-Cell Development Mouse | CHAPTER 10 331 Human Placenta Yolk sac Fetal liver AGM Specification Mesoderm Emergence Maturation Pre-HSC Expansion HSC Quiescence/ self-renewal HSC HSC Mouse 7.5 10.5 12.5 15.5 days 21 28 40 70 days Birth Human Birth Placenta Fetal liver AGM Primitive streak Bone marrow Yolk sac FIGURE 10-2 The anatomy and timing of the earliest become blood cells emerge first as rapidly proliferating pre-HSCs and eventually mature into relatively quiescent hematopoietic stem cells that populate the bone marrow. The colored bars in the timeline illustrate the ages at which the various murine and human hematopoietic sites are active. Mesoderm (gray); generation of fetal HSC (yellow); active hematopoietic differentiation (red); emergence of functional, adult-type HSCs (blue). [Adapted from H. K. Mikkola and S. H. stages in hematopoiesis. (a) Blood-cell precursors are initially found in the yolk sac (yellow), then spread to the placenta fetal liver (pink), and aorta-gonad-mesonephros (AGM) region (green), before finding their adult home in the bone marrow. The mouse embryo is shown at 11 days of gestation; the human embryo at the equivalent 5 weeks of gestation. (b) In the embryo, cells within the primitive streak mesodermal tissue adopt either hematopoietic (blood-cell forming) or vascular (blood-vessel forming) fates. Those destined to Orkin, 2006, The journey of developing hematopoietic stem cells, Development 133:3733–3744, Figures 1 and 2. ] development, the whole process of blood-cell generation must shift location several times before moving into its final home. The gestation period for mice is 19 to 21 days. Hematopoiesis begins, in the mouse, around 7 days post fertilization (Figure 10-2) when precursor cells in the yolk sac begin differentiating to form primitive, nucleated, erythroid cells that carry the oxygen the embryo needs for early development. Fetal HSCs capable of generating all blood-cell types can be detected in the early aorta-gonad-mesonephros (AGM) region on day 8, when the fetal heart starts beating. On day 10, mature HSCs capable of completely repopulating the hematopoietic system of irradiated adult mice can be isolated from the AGM, and by day 11 they can be found in the yolk sac, placenta, and fetal liver. Between days 11.5 and 12.5, there is rapid expansion of the placental HSC pool, at the end of which time the placenta holds more HSCs than either the AGM or the yolk sac. By day 13.5, the number of HSCs in the placenta begins to decrease while the HSC pool in the liver continues to expand. As the mouse embryo completes its development, the predominant site of HSC generation remains within the fetal liver, but some hematopoiesis can be detected in the spleen in the perinatal (around the time of birth) period. The number of fetal liver HSCs in mice reaches a maximum of approximately 1000 by days 15.5 to 16.5 of embryonic development, after which it starts to decline. Within the fetal liver, HSCs differentiate to form progenitor cells. At the 332 PA R T I V | Adaptive Immunity: Development earliest time points, hematopoiesis in the fetal liver is dominated by erythroid progenitors that give rise to the true, enucleated mature erythrocytes, in order to ensure a steady oxygen supply to the growing embryo, and myeloid and lymphoid progenitors gradually emerge. Pre-B cells (precursor-B cells), defined as cells that express immunoglobulin in their cytoplasm but not on their surfaces, are first observed at day 13 of gestation, and surface IgM-positive B cells are present in detectable numbers by day 17. HSCs first seed the bone marrow at approximately day 15, and over a period of a few weeks, the bone marrow takes over as the main site of B-cell development, remaining so throughout post-natal life. Hematopoiesis in the Fetal Liver Differs from That in the Adult Bone Marrow Developing B cells in the fetal liver differ in important ways from their counterparts in adult bone marrow. The liver is the primary site of B-cell generation in the fetus, and provides the neonatal animal with the cells it needs to populate its nascent immune system. In order to accomplish this, hematopoietic stem cells and their progeny must undergo a phase of rapid proliferation, and fetal liver HSCs, as well as their daughter cells, undergo several rounds of cell division over a short time. In contrast, HSCs derived from the bone marrow of a healthy adult animal are relatively quiescent. B cells generated from fetal liver precursors are predominantly B-1 B cells, which will be described more fully in Chapter 12. Briefly, B-1 B cells are primarily located in the body (specifically the peritoneal and pleural) cavities. They are therefore well-positioned to protect the gut and the lungs, which are the major ports of entry of microbes in the fetus and neonate. Antibodies secreted by B-1 B cells are broadly cross-reactive; many bind to carbohydrate antigens expressed by a number of microbial species. Since terminal deoxynucleotidyl transferase (TdT) is minimally expressed at this point in ontogeny, and the RAG1/2 recombinase proteins appear not to use the full range of V, D, and J region gene segments at this stage in embryonic development, the immunoglobulin receptors of B-1 B cells express minimal receptor diversity. In expressing an oligoclonal (few, as opposed to many, clones) repertoire of B-cell receptors that bind to a limited number of carbohydrate antigens shared among many microbes, B-1 B cells occupy a functional niche that bridges the innate and adaptive immune systems. We will describe B-1 B-cell development further at the end of this chapter. Over a period of 2 to 4 weeks after birth, the process of hematopoiesis in mice shifts from the fetal liver and spleen to the bone marrow, where it continues throughout adulthood. The B-1 B-cell population represents an exception to this general rule, as it is self-renewing in the periphery. This means that new daughter B-1 B cells are generated continually from preexisting B-1 B cells in the peritoneal and pleural cavities, and in those other parts of the body in which B-1 B cells reside. These daughter cells use the same receptors as their parents, and no new V(D)J recombinase activity is required. In humans, the sequence of events is similar to that described for the mouse, but the time frame is obviously somewhat elongated. Blood-cell precursors first appear in the yolk sac in the third week of embryonic development, but these cells, like their analogues in the mouse provide primarily erythroid progenitors and are not capable of generating all subsets of blood cells. The first cells capable of entirely repopulating an adult human hematopoietic system arise in the AGM region of the embryo and/or the yolk sac. By the third month of pregnancy, these HSCs migrate to the fetal liver, which then becomes responsible for the majority of hematopoiesis in the fetus. By the fourth month of pregnancy, HSCs migrate to the bone marrow, which gradually assumes the hematopoietic role from the fetal liver until, by the time of birth, it is the primary generative organ for blood cells. Prior to puberty in humans, most of the bones of the skeleton are hematopoietically active, but by the age of 18 years only the vertebrae, ribs, sternum, skull, pelvis, and parts of the humerus and femur retain hematopoietic potential. Just as B-cell development in the fetus and neonate differs from that in the adult, so does B-cell hematopoiesis in the aging animal. Clinical Focus Box 10-1 describes some aspects of B-cell development that alter as humans age. B-Cell Development in the Bone Marrow In Chapter 2 (Figure 2-5), we presented the structure of bone and bone marrow. The bone marrow microenvironment is a complex, three-dimensional structure with distinctive cellular niches which are specialized to influence the development of the cell populations that mature there. A dense network of fenestrated (leaky) thin-walled blood vessels— the bone marrow sinusoids—permeates the marrow, allowing the passage of newly formed blood cells to the periphery and facilitating blood circulation through the marrow. In addition to serving as a source of hematopoietic stem cells, bone marrow also contains stem cells that can differentiate into adipocytes (fat cells), chondrocytes (cartilage cells), osteocytes (bone cells), myocytes (muscle cells), and potentially other types of cells as well. Each of these different classes of stem cells requires specific sets of factors, secreted by particular bone marrow stromal cells to enable their proper differentiation. What are bone marrow stromal cells? The term stroma derives from the Greek for mattress, and a stromal cell is a general term that describes a large adherent cell that supports the growth of other cells. During B-cell development, bone marrow stromal cells fulfill two functions. First, by interacting with adhesion molecules on the surfaces of HSCs and progenitor cells, stromal cells retain the developing cell populations in the specific bone marrow niches where they can receive the appropriate molecular signals required for B-Cell Development | CHAPTER 10 333 Box 10-1 CLINICAL FOCUS B-Cell Development in the Aging Individual People of retirement age and older represent a greater segment of the population than they used to, and these older individuals expect to remain active and productive members of society. However, physicians and immunologists have long known that the elderly are more susceptible to infection than are young men and women, and that vaccinations are less effective in older individuals. In this feature, we explore the differences in B-cell development between younger and older vertebrates, which may account for some of these immunological disparities between adult and older individuals. Aging individuals display deficiencies in many aspects of B-cell function, including a poor antibody response to vaccination, inefficient generation of memory B cells, and an increase in the expression of autoimmune disorders. Does this reflect defective functioning only in the mature antigen-responsive B-cell population, or does it result from problems manifested during earlier stages in B-cell development? Current research demonstrates that aging individuals display a range of shortcomings in developing B cells. Experiments employing reciprocal bone marrow chimeras—in which aging HSCs were transplanted into young recipients or HSCs from young mice were injected into aging recipients—have shown that the suboptimal process of B-cell development in aging individuals results from deficiencies in both the aging stem cells and in the supporting stromal cells. For example, bone marrow stromal cells from aging mice secrete lower levels of IL-7 than do stromal cells from younger animals, suggesting an environmental defect in the aging bone marrow. However, study of isolated, aging B-cell progenitors reveals that they also respond less efficiently to IL-7 than do B cells from younger mice, and so the IL-7 response in aging individuals is affected at both the secretory and recipient-cell levels. Indeed, the problems encountered by developing B cells from aging indi- viduals start at the very beginning of their developmental program. The epigenetic regulation of HSC genes in aging mice is compromised, resulting in diminished levels of HSC self-renewal. Furthermore, the balance between the production of myeloid versus lymphoid progenitors is shifted in older individuals, with down-regulation of genes associated with lymphoid specification and a correspondingly enhanced expression of genes specifying myeloid development. The net effect of these changes in the HSC population is a reduction with age in the numbers of early B-cell progenitors, which is reflected in a decrease in the numbers of pro- and pre-B-cell precursors at all stages of development. Detailed studies of the expression of particular genes important in B-cell development demonstrate that the expression of important transcription factors, such as those encoded by the E2A gene, is reduced in older animals. Furthermore, the Rag genes, as well as the gene encoding the surrogate light-chain component, 5, are down-regulated in older animals compared with young adults, resulting in a reduction in the bone marrow output of immature B cells. Multiple mechanisms therefore help to explain why the numbers of B cells released from the bone marrow are smaller in aging than in younger individuals. But is the antigen recognition capacity— the quality—as well as the quantity of B cells different between the two populations? In particular, do B cells from aging mice express a repertoire of receptors similar to those obtained from younger animals? The answer to this question has come from the development of techniques that enable a global assessment of repertoire diversity. Study of the sizes and sequences of CDR3 regions from large numbers of human B cells suggests that in aging individuals the size of the repertoire (the number of different B-cell receptors an individual expresses) is drastically diminished, and that this decrease in repertoire diversity correlates with a reduction in the health of the aging patient. The mechanisms for this age-related repertoire truncation appear to be complex. A decrease in output of immature B2 cells from the bone marrow could provide the opportunity for B-1 B cells to increase their share of the peripheral B-cell niche, and as is well appreciated, B-1 B cells have a less diverse receptor repertoire than do B-2 B cells. A lifetime of generating memory cells may also result in an individual having less room in B-cell follicles for newly formed B cells to enter, and a decreased concentration of the homeostatic regulatory cytokines may make it more difficult for primary B cells to compete with their more robust memory counterparts. Clearly, this is an area of increasing clinical interest as the average age of the population of the developed world continues to increase. REFERENCES Cancro, M. P., et al. 2009. B cells and aging: Molecules and mechanisms. Trends in Immunology 30:313–318. Dorshkin, K., E. Montecino-Rodriguez, and R. A. Signer. 2009. The ageing immune system: Is it ever too old to become young again? Nature Reviews Immunology 9:57–62. Goodnow, C. C. 1992. Transgenic mice and analysis of B-cell tolerance. Annual Reviews Immunology 10:489–518. Labrie, J. E., 3rd, A. P. Sah, D. M. Allman, M. P. Cancro, and R. M. Gerstein. 2004. Bone marrow microenvironmental changes underlie reduced RAG-mediated recombination and B cell generation in aged mice. Journal of Experimental Medicine 200:411–423. Nemazee, D. A., and K. Bürki. 1989, February. Clonal deletion of B lymphocytes in a transgenic mouse bearing anti-MHC class I antibody genes. Nature 337:562–566. doi:10.1038/337562a0 Van der Put, E., E. M. Sherwood, B. B. Blomberg, and R. L. Riley. 2003. Aged mice exhibit distinct B cell precursor phenotypes differing in activation, proliferation and apoptosis. Experimental Gerontology 38:1137–1147. 334 PA R T I V | Adaptive Immunity: Development Endothelial cell Osteoblast Bone IL-7 expressing cell Pro-B cell Pre-proB cell Pre-B cell Immature B cell HSC Medullary vascular sinus FIGURE 10-3 HSCs and B-cell progenitors make contact with different sets of bone marrow cells as they progress through their developmental program. HSCs begin their developmental program close to the osteoblasts (top). An HSC is also shown entering from the blood (left-hand side), illustrating the fact that HSCs are capable of recirculation in the adult animal. Progenitor cells then move to gain contact with CXCL12-expressing stromal cells, where they mature into pre-pro B cells. By the time differentiatheir further differentiation. Second, diverse populations of stromal cells express different cytokines. At various points in their development, progenitor and precursor B cells must interact with stromal cells secreting particular cytokines, and thus the developing B cells move in an orderly progression from location to location within the bone marrow. This progression is guided by chemokines secreted by particular stromal cell populations. For example, HSCs begin their life in close contact with osteoblasts located close to the lining of the endosteal (bone marrow) cavity. Once differentiated to the pre-pro-B-cell stage, the developing B cells require signals from the chemokine CXCL12, which is secreted by a specialized set of stromal cells, in order to progress to the pro-B-cell stage. Pro-B cells then require signaling from the cytokine IL-7, which is secreted by yet another stromal cell subset (Figure 10-3). Many of these stromal cell factors serve to induce the expression of specialized transcription factors important in B-cell development. The Stages of Hematopoiesis Are Defined by Cell-Surface Markers, Transcription-Factor Expression, and Immunoglobulin Gene Rearrangements Full characterization of a developmental pathway requires that scientists understand the phenotypic and functional IL-7 CXCL12 Plasma cell CXCL12 reticular cell tion has progressed to the pro-B-cell stage, the developing cell has moved to receive signals from IL-7-producing stromal cells. After leaving the IL-7-expressing stromal cell, the pre-B cell completes its differentiation and leaves the bone marrow as an immature B cell. CXCL-12 is shown in purple; IL-7 in blue. [Adapted from T. Nagasawa, 2006, February, Microenvironmental niches in the bone marrow required for B-cell development, Nature Reviews Immunology 6:107–116. doi:10.1038/ nri1780] characteristics of each cell type in that pathway, as well as the molecular signals and transcription factors that drive differentiation at each stage. Cells at particular stages of differentiation can be characterized by their surface molecules, which include cell-surface antigens, adhesion molecules, and receptors for chemokines and cytokines. They are also defined by the array of active transcription factors that determine which genes are expressed at each step in the developmental process. Finally, in the case of B cells, the developmental stages are defined by the status of the rearranging heavy- and light-chain immunoglobulin genes. B-cell development is not yet completely understood; however, most of the important cellular intermediates have been defined, and developmental immunologists are gradually filling in the gaps in our knowledge. Investigators delineating the path of B-lymphocyte differentiation employed three general experimental strategies. First, they generated antibodies against molecules (antigens or markers) present on the surface of bone marrow cells. They then determined which of these molecules were present at the same time as other antigens, and which combinations of antigens appeared to define unique cell types (Figure 10-4a). In addition, culturing cells in vitro that bear known cellsurface antigens, followed by flow cytometric analysis of the daughter cells generated in culture, enabled them to describe the sequential expression of particular combinations of cellsurface molecules (Figure 10-4b). B-Cell Development | CHAPTER 10 335 (d) (a) Knocking out particular transcription factors (TFs) stops development at particular points. Characterization of progenitors bearing different sets of cell surface molecules. Knocking out TF1 leaves only this population. Therefore TF1 is required to progress to VHD recombination. (b) Determining sequence of marker expression by culturing cells from each stage. Culturing cells with the red antigen gives rise to daughter cells of both types. Culturing cells with only the blue antigen gives rise to the cells bearing both blue and green antigens, but never cells bearing the red antigens. Therefore, we can sequence the three cell types in this way. Knocking out TF2 leaves these two populations. Therefore TF2 is required for progression to light chain rearrangement. (e) (c) Sequencing of different antigen-bearing cells by analyzing each population for the stage of V(D)J rearrangements in heavy and light chains. D-JH recombination only VHDJH recombination completed VH and VL recombination completed Placing GFP under the control of the TF2 promoter reveals that TF2 expression occurs in these two cell populations. Clearly, it is turned on during the end of the blue stage, and is needed for progression to the blue and green stage of development. VHDJH recombination completed VH and VL recombination completed FIGURE 10-4 Experimental approaches to the staging and characterization of B-cell progenitors. In this figure, the different icons do not represent specific antigens, but are used in order to illustrate the principle of the experiment. Investigators delineated the stage of B-cell development by using flow cytometry to characterize the cell-surface expression of developmental markers and molecular biology to correlate the expression of specific markers with the stage of immunoglobulin gene rearrangement. The requirement for transcription factor activity at each step was determined using both knockout and knockin genetic approaches. (See text for details.) Second, by sorting cells bearing particular combinations of cell-surface markers, and analyzing those cell populations for the occurrence of immunoglobulin gene rearrangements, scientists were able to confirm the staging of the appearance of particular developmental antigens. For example, an antigen that appears on a cell in which no variable region gene rearrangement has taken place is clearly expressed very early in B-cell differentiation. Similarly, an antigen present on the surface of a cell that has rearranged heavychain, but not light-chain, genes defines a stage in B-cell differentiation later than that defined by the marker described above, whereas an antigen present on a cell that has undergone both heavy-chain and light-chain rear- rangement characterizes a very late stage in B-cell development. In this way, cell-surface markers were defined that could serve as indicators of particular steps in B-cell differentiation (Figure 10-4c). Third, investigators use the power of knockout genetics to determine the effects on B-cell development of eliminating the expression of particular genes, such as those encoding particular transcription factors (Figure 10-4d). For example, knocking out a gene encoding a particular transcription factor eliminates an animal’s ability to complete VH to DJH recombination while still allowing D to JH rearrangement. This tells us that the particular transcription factor in question is not necessary for stages of B-cell differentiation 336 PA R T I V | Adaptive Immunity: Development ADVANCES The Role of miRNAs in the Control of B-Cell Development Geneticists have long known that only a small fraction of chromosomal DNA specifies protein sequences, and early papers relegated the nonprotein-coding DNA segments to the somewhat ignominiously described status of “junk DNA.” In 1993, however, scientists studying the genome of the nematode C. elegans described groundbreaking investigations of some of the nonprotein-coding sequences that they had identified as having been transcribed but not translated. They showed that these primary transcripts were processed into small pieces of RNA, 18 to 30 nucleotides in length, that were capable of exerting control over the level of expression of mRNA. The biosynthesis of these micro-RNAs follows a similar form in eukaryotes as diverse as C. elegans and humans (Figure 1). Fully capped and polyadenylated RNAs (pri-microRNAs) are synthesized by RNA polymerase II and are then cleaved into a hairpin-shaped 70- to 100-nucleotide premicroRNA by the nuclear RNAase Drosha, which works in tandem with a second double-stranded RNA-binding protein, DGCR8. The cleaved pre-micro-RNA is then exported to the cytoplasm, where a second RNAase, Dicer, acting in association with two other proteins, processes it to an 18- to 30-nucleotide miRNA duplex, consisting of the mature miRNA and its anti-sense strand. In a final step, the mature miRNA, now single stranded, associates with a protein complex called the RNA-induced silencing complex, or RISC. Nucleus DNA Cytoplasm RNA polymerase mRNA-target pri-microRNA Drosha/ DGCR8 RISC pre-microRNA RISC Dicer PACT and TRBP FIGURE 1 The generation of functioning microRNAs (miRNAs). Just like mRNA, miRNA species are transcribed as long, capped and polyadenylated RNA species (pri-miRNA) by RNA polymerase ll. They are then cleaved by a nuclear RNase, Drosha, into a hairpin shaped nucleotide precursor molecule, termed a pre-miRNA. Drosha works in a protein complex with the protein DGCR8 (DiGeorge syndrome chromosomal region 8). Pre-miRNAs are then exported to the cytoplasm where a second ribonuclease, Dicer, in association with the proteins PACT and TRBP processes the pre-miRNA into a 19 to 24 nucleotide miRNA duplex, by removing the terminal loop. Next, a protein complex called RISC (RNA-Induced Silencing Complex) binds to one of the two strands of the duplex. The strand of miRNA that binds to RISC is the mature miRNA, and it drives the RISC enzyme to the target mRNA, resulting in mRNA silencing and/or destruction. [Adapted from Vasilatou et al., 2009. The role of microRNAs in normal and malignant hematopoiesis. European Journal of Hematology, 84, 1 to 16. Figure 1.] The mature miRNA operates by complementary binding of a so-called “seed” region of 6 to 8 nucleotides at its 5 end to a region on its target mRNA. Once the miRNA has bound, three things can happen: the target mRNA can be directly tar- prior to D to JH rearrangement, but it is required for one or more stages, starting with VH to DJH recombination. One drawback of the knockout approach, however, is that it only defines the first stage in differentiation at which the transcription factor is required. More recent variations have exploited knockin genetics (Figure 10-4e) to generate animals that express fluorescent markers under the control of geted for cleavage; the mRNA can be destabilized; or translation from the mRNA can be repressed. Furthermore, we now know that a single miRNA can target the synthesis of many proteins, and each mRNA can be the target of more than one the transcription factor promoters, so that every point at which the transcription factor is expressed can be delineated. The staging of transcription factor expression is then correlated with the expression of both cell-surface markers and immunoglobulin gene rearrangement. The sequence of B-cell development described in the next few sections has been elucidated using a combination of these strategies. B-Cell Development | CHAPTER 10 337 BOX 10-2 miRNA, thus adding to both the flexibility and the complexity of this mode of control over gene expression. But does regulation by miRNAs operate during B-cell development? Several lines of evidence suggest that the answer to this question is an unequivocal “yes.” From a theoretical standpoint, it is clear that the developmental changes that occur as B cells mature require rapid changes in the concentrations of such important proteins as transcription factors and pro- and anti-apoptotic molecules, among other regulatory proteins. The need for such rapid alterations in protein concentrations can be met efficiently by the type of post-transcriptional control mechanisms mediated by miRNAs. Conditional loss of the gene encoding the Dicer nuclease destroys all capacity to synthesize mature miRNAs. Ablation of Dicer in early B-cell progenitors resulted in a developmental block at the pro- to pre-Bcell transition. In these experiments, the pro-apoptotic molecule Bim, was expressed at higher concentrations in Dicer-ablated than in normal B-cell progenitors. Sequences in the 3 untranslated region of the Bim gene were found to be complementary to miRNAs of the 17~92 family, suggesting that members of this family normally down-regulate Bim at this stage of development, enabling B cells to pass through this transition. The 17~92 family of miRNAs was also shown to affect expression of TdT and hence N-sequence addition. Other alterations in immunoglobulin gene expression were also observed in the absence of 17~92 miRNAs, including an increase in the expression of sterile transcripts. These data collectively demonstrate that miRNAs are important in the control of the pro- to pre-B-cell transition and affect both the expression of pro-apoptotic molecules and the nature of the Ig repertoire. As one would predict from these collective results, animals with increased expression of miR-17~92 family members express lower levels of Bim and suffer from a lympho-proliferative disorder and autoimmune disease. Members of the 17~92 miRNA family were also found to control the levels of the Pten protein, which acts as an inhibitor of the pro-survival PI3 kinase and Akt signaling pathway. An increase in the levels of the miR-17~92 molecules allows for greater destruction of the Pten mRNA, resulting in increased cell survival and a corresponding increase in the number of lymphocytes available to proliferate. Other investigators have addressed the question of which miRNAs are expressed at different stages of B-cell development. A combination of genomic (in silico analyses) and more classical molecular biological approaches have identified several miRNA species and/or families that are implicated in the control of B-cell development. Perhaps the most well studied miRNA is miR-150, which is highly expressed in mature and resting B cells but not in their progenitors. The miR-150 molecule has been shown to depress the level of expression of the transcription factor c-Myb, known to be essential for the control of B-cell development. As might have been predicted, the pattern of c-Myb expression in lymphocyte development is complementary to that of miR-150, in that c-Myb is highly expressed in lymphoid progenitors and is down-regulated upon their maturation; in addition, transcriptional analysis confirmed that miR-150 is Recently, attention has begun to focus on small molecular weight miRNA species, which have profound effects on the stability of mRNA and hence on the expression of particular proteins. In Advances Box 10-2, we describe the effects of some of the miRNAs that have recently been shown to affect B-cell differentiation. This is currently an extremely active research area. an important factor in the regulation of the levels of the c-Myb transcription factor during B-cell development. MiR-150 is also implicated in B-1/B-2 lineage specification. B-cell-specific deletion of the c-myb gene stops B-cell development at the pro- to pre-B transition and also leads to the complete disappearance of the B-1 subset of B cells. If miR-150 is responsible for down-modulating the levels of c-Myb in vivo, then it might be predicted that a deficiency in miR-150 would result in an opposite phenotype to that expressed by a c-Myb deficient animal, and such was indeed found to be the case. Deficiency of miR-150 was found to result in an expansion of the B-1-B-cell pool, with a resulting increase in the levels of IgM antibody secretion. The study of miRNA control of mammalian gene expression and lymphocyte development is still in its infancy, but the ability to manipulate and isolate cells at discrete stages in the developmental sequence provides a particularly tractable system in which to analyze the range of actions of different families of miRNAs. This field will undoubtedly be one to watch. REFERENCES Baltimore et al., 2008. MicroRNAs: new regulators of immune cell development and function. Nature Immunology 9:839–845. Koralov, S. B., et al. 2008. Dicer ablation affects antibody diversity and cell survival in the B lymphocyte lineage. Cell 132:860–874. Vasilatou, D., S. Papageorgiou, V. Pappa, E. Papageorgious, and J. Dervenoulas. 2009. The role of microRNAs in normal and malignant hematopoiesis. European Journal of Haematology 84:1–16. Xiao, C., and K. Rajewsky. 2009. MicroRNA control in the immune system: Basic principles. Cell 136:26–36. The Earliest Steps in Lymphocyte Differentiation Culminate in the Generation of a Common Lymphoid Progenitor In this section, we will describe the process by which an HSC in the bone marrow develops into a CLP. Unless otherwise specified, the developmental pathway we describe refers to 338 PA R T I V | Erythrocyte Megakaryocyte Monocyte T-cell progenitor Natural killer (NK) cell Dendritic cell (DC) Adaptive Immunity: Development c-Kit sca-1 flt-3 CD34 IL-7R Rag1/2 and TDT + + – – – – Multipotent progenitor (MPP) + + – + – – Lymphoid-primed multipotent progenitor cell (LMPP) + + + – – – Early lymphoid progenitor cell (ELP) + + + – +/– + low low + – + + Hematopoietic stem cell (HSC) Common lymphoid precursor (CLP) Pre-pro B cell FIGURE 10-5 Expression of cell-surface markers on HSC and lymphoid progenitor cells. The maturation of HSCs into lymphoid progenitors, and the progressive loss of the ability to differentiate into other blood-cell lineages can be followed by the expression of the cellsurface markers as well as by the acquisition of RAG and TdT activity. that followed by the predominant B-2-B-cell population. Specific aspects of development that differ among the various B-cell subsets will be addressed toward the end of this chapter. HSCs HSCs are both self-renewing (they can divide to create identical copies of the parent cell) and multipotential (they can divide to form daughter cells that are more differentiated than the parent cell and that can develop along distinct blood-cell lineages), and can give rise to all cells of the blood. HSCs maintain a relatively large number of genes in a socalled “primed” state, and an individual HSC may possess primed genes characteristic of multiple cell lineages. Primed chromatin is associated with lower-than-usual numbers of nucleosomes, is more accessible to enzyme activity than the majority of chromatin in the cell, and shows histone methylation and acetylation patterns characteristic of active chromatin. Depending on the environmental stimuli to which any given HSC is exposed, transcription factors may drive the cell down a number of possible developmental pathways. During the differentiation process that follows, primed chromatin regions containing genes that are not needed for the selected developmental pathway are shut down. In HSCs bound for a B-cell fate, the transcription factors Ikaros, Purine box factor 1 (PU.1), and E2A participate in the earliest stages of B-lineage development. Ikaros recruits chromatin remodeling complexes to particular regions in the DNA and ensures the accessibility of genes necessary for B-cell development. PU.1 presides over a leukocytic “balancing act”; low levels of PU.1 favor lymphoid differentiation, whereas cells expressing higher levels of PU.1 veer off to a myeloid fate. The level of PU.1 protein expressed is in turn regulated by the transcriptional repressor Gfi1, which downregulates the expression of PU.1 to the levels necessary for progression down the B-cell pathway. E2A expression contributes to the maintenance of the HSC pool by participating in the regulation of cell cycle control in this population. As noted in Chapter 9, HSCs express the cell-surface molecule c-Kit (CD117), which is the receptor for stem cell factor (SCF). SCF is a cytokine that exists in both membrane-bound and soluble forms and the SCF-c-Kit interaction is critical for the development, in adult animals, of multipotential progenitor cells (MPPs). Membrane-bound SCF plays a role in retaining the HSCs and its daughter progenitor cells in the appropriate environmental niches in the bone marrow. HSCs also express the stem cell associated antigen-1 (Sca-1). Both c-Kit and Sca-1 are expressed in parallel on early progenitor cells, and the levels of their expression drop as the cells commit to a particular cell lineage (Figure 10-5). HSCs are often described as being Lin, a designation that refers to the fact that they have no “lineage markers” characteristic of a particular blood-cell subpopulation. B-Cell Development MPPs MPPs generated on receipt of SCF/c-Kit signaling lose the capacity for extensive self-renewal, but retain the potential to differentiate into several different hematopoietic lineages. MPP cells retain the expression of c-Kit and Sca-1 and transiently express the molecule CD34. Indeed, antibodies to CD34 are used clinically to isolate cells at this stage in hematopoiesis. MPPs also express the chemokine receptor CXCR4, which enables them to bind the stromal cell-derived chemokine CXCL12. The interaction between CXCL12 and CXCR4 is important in ensuring that the progenitor cell occupies the correct niche within the bone marrow (see Figure 10-3). LMPPs The progenitor cell on its way to becoming a B cell then begins to express the fms-related tyrosine kinase 3 receptor (flt-3). Flt-3 binds to the membrane-bound flt-3 ligand on bone marrow stromal cells and signals the progenitor cell to begin synthesizing the IL-7 receptor (IL-7R, CD127). Flt-3 is expressed on B-cell progenitors from this point until the pro-B stage and acts synergistically with IL-7R to promote the growth of cells bearing flt-3 and IL-7R. The expression of flt-3 on the surface of the developing cell marks the loss of the potential of the MPP cell to develop into erythrocytes or megakaryocytes, and therefore characterizes a new level of cell commitment; however, this progenitor still retains the capacity to develop along either the myeloid or the lymphoid pathways. These cells, now c-Kit, Sca-1, and flt-3, are termed lymphoid-primed, multipotential progenitors (LMPPs) (see Figure 10-5). As they become further committed to the lymphoid lineage, levels of the stem-cell antigens c-Kit and Sca-1 fall, and cells destined to become lymphocytes begin to express RAG1/2 and terminal deoxynucleotidyl transferase (TdT) (Chapter 7). Expression of the genes encoding RAG1/2, TdT, IL-7R, and the B-cell-specific transcription factor EBF1 are all up-regulated at the end of this stage. ELPs Expression of RAG1/2 defines the cell as an early lymphoid progenitor cell (ELP). A subset of ELPs migrates out of the bone marrow to seed the thymus and serve as the T-cell progenitors (discussed in Chapter 9). The rest of the ELPs remain in the bone marrow as B-cell progenitors. On these cells, the levels of the early c-Kit and Sca-1 antigens decrease as the levels of the IL-7R increase, and the ELP now develops into a CLP. CLPs At the CLP stage, the progenitor on its way to B-cell commitment still retains the potential to mature along the NK, conventional DC, or T-cell lineages. At this point in development, signals received through IL-7R promote cell survival and enhance the production of EBF-1 and other transcription factors that are required for later steps | CHAPTER 10 339 in the B-cell differentiation pathway. Signaling through the IL-7R occurs via pathways familiar from Chapter 4. Specifically, an IL-7R-mediated JAK-STAT pathway induces the up-regulation of the anti-apoptotic molecule Mcl1. Signaling through IL-7R also results in the upregulation of the C-myc and N-myc genes, which signal the cell proliferation characteristic of the later, pro-B-cell stage. CLPs are c-Kitlow, Sca-1low, and IL-7R and have lost myeloid potential. However, as a CLP destined to differentiate along the B cell pathway matures, the chromatin containing the immunoglobulin locus becomes increasingly accessible and the developing lymphocyte approaches the point at which it is irrevocably committed to the B-cell lineage. The Later Steps of B-Cell Development Result in Commitment to the B-Cell Phenotype Figure 10-6 illustrates the expression of cell-surface markers, and the patterns of rearrangement of immunoglobulin heavy-chain and light-chain genes starting at the pre-pro B-cell stage of development. The stages of B-cell differentiation have been defined by more than one group of scientists and, as a result, two systems of nomenclature are in common use. The first, and most widely used, is the Basel nomenclature (pre-pro, pro, pre-B, immature B) developed by Melchers and colleagues. The second (A, B, C, C, D, E) is that defined by Hardy et al., and the process by which this system of classification of B-cell development was established is described in detail in Classic Experiment Box 10-3. Pre-Pro B Cells With the acquisition of the B-cell lineage-specific marker B220 (CD45R), and the expression of increasing levels of the transcription factor EBF1, the developing cell enters the prepro-B-cell stage. EBF1 is an important transcription factor in lymphoid development, and therefore transcription of the Ebf1 gene is itself under the control of multiple transcription factors (Figure 10-7). These each bind at distinct promoter regions, and hence the level of transcription of the Ebf1 gene can vary considerably depending on the combination of controlling factors present at any particular developmental stage. At the pre-pro-B-cell stage, EBF-1, along with E2A, binds to the immunoglobulin gene, promoting accessibility of the D-JH locus and preparing the cells for the first step of Ig gene recombination. EBF-1 is also essential to the full expression of many B-cell proteins, including Ig,Ig (CD79,), and the genes encoding the pre-B-cell receptor, which will be expressed when heavy-chain VDJ recombination is complete. Pre-pro B cells remain in contact with CXCL-12-secreting stromal cells in the bone marrow. However, the onset of D to JH gene recombination classifies the cell as an early pro-B cell, and at this stage the developing cell moves within the bone marrow, seeking contact with IL-7 secreting stromal cells. PA R T I V 340 | Adaptive Immunity: Development Status of Ig genes Surface Ig B220 receptor c-Kit IL-7R CD25 CD19 (CD45R) expression Pre-pro (Fr. A)1 GL2 None – lo – – + Early pro (Fr. B) DJH None lo lo/+ – + + Late pro (Fr. C) Some VHDJH None lo + +/– + + Large Pre (Fr. C') VHDJH Pre-BCR – + + + + Small Pre (Fr. D) VHDJH VL JL rearrangement begins Decreasing levels of Pre-BCR – + + + + IgM – + + + + Cell stage 1st checkpoint Immature B VHDJH (Fr. E) VL JL FIGURE 10-6 Immunoglobulin gene rearrangements and expression of marker proteins during B-cell development. The expression of selected marker proteins is correlated with the extent of Ig gene rearrangement during B-cell development from the pre-pro B cell to the immature T1 B-cell stage. (See text for details.) [Adapted from K. Samitas, J. Lötvall, and A. Bossios, B cells: From early development to regulating allergic diseases, Archivum Immunologiae et Therapiae Experimentalis 58:209–225, Figure 1.] 2nd checkpoint Pre-BCR BCR 1 Labeled fractions refer to the “Hardy nomenclature,” described in the Classic Experiment Box 10-1. 2 GL = germ line arrangement of heavy and/or light chain V region segments IL-7 IL-7Rγ IL-7Rα FIGURE 10-7 The interplay of transcripCytoplasm STAT5 P P STAT5 • Transcription of B-cell associated genes survival • B-cell specification and commitment survival E2A N-myc C-myc EBF1 B-cell proliferation PAX5 Nucleus D-JH recombination VHDJH recombination Notch1 tion factors during early B-cell development. Dimerization and activation of the transcription factor STAT5 is stimulated by IL-7 binding to its receptor. STAT5 stimulates B-cell proliferation by activating the proliferative control proteins N-myc and C-myc. STAT5 collaborates with E2A proteins to promote the expression of early B-cell factor 1 (EBF1). EBF1 in turn promotes the expression of PAX5, and together the E2A proteins, EBF1, and PAX5 activate many genes leading to B-cell lineage specification and commitment. PAX5 and EBF1 both participate in positive feedback loops that enhance the levels of both EBF1 and PAX5 transcription. [Adapted from B. L. Kee, 2009, E and ID proteins branch out, Nature Reviews Immunology 9:175–184.] B-Cell Development | CHAPTER 10 341 BOX 10-3 CLASSIC EXPERIMENT The Stages of B-Cell Development: Characterization of the Hardy Fractions Richard Hardy’s laboratory was one of the first to combine flow cytometry and molecular biology in experiments designed to analyze lymphocyte maturation. In this feature, we describe what those researchers did and how they generated a model of the sequencing of the stages of B-cell development from their data. When Hardy and colleagues began their characterization of B-cell lineage development in the early 1990s, prior work using molecular analysis of long term bone marrow cell lines had already established the sequential rearrangement of heavychain and light-chain immunoglobulin genes. In addition, the expression of a number of cell-surface markers on bone marrow cells had been measured, and several of these antigens had been shown to be co-expressed with B220 (CD45R), which had already been established as a B-cell differentiation antigen. Hardy’s approach was to characterize the sequence of expression of those antigenic markers that were found on the same cells as B220. The hypothesis was that some of these markers may be expressed on early B-cell progenitors and might therefore help to generate a scheme of B-cell development. In order to place cells expressing different combinations of markers into a developmental lineage, Hardy then sorted cells bearing each combination of his selected markers, and placed them into co-cultures with a bone marrow stromal cell line. After defined times in culture, he harvested the hematopoietic cells and re-characterized their surface marker expression. The markers used in these experiments included B220 (CD45R) and CD43 (leukosialin), which had previously been shown to be expressed on granulocytes and all T cells, but was not present on mature B cells, with the exception of plasma cells. In addition, their experiments employed antibodies directed against Heat Stable Antigen, or HSA (CD24) and BP-1, an antigen on bone Start with bone marrow cells Sort for B220+CD43+ Analyzed B220+ CD43+ cells for HSA and BP-1 expression HSABP-1Fraction A Pre-pro-B cells HSA+ BP-1Fraction B Early pro-B cells HSA+ BP-1+ Fraction C Lower levels of HSA Fraction C Late pro-B cells Higher levels of HSA Fraction C' Large (early) pre-B cells FIGURE 1 The isolation of Hardy’s fractions. A, B, C, C. Bone marrow cells were sorted for cells bearing B220 and CD43 and then analyzed for their expression of the cell-surface markers HSA and BP-1. marrow cells. Both HSA and BP-1 had been previously shown to be differentially expressed at varying stages of lymphoid differentiation. The first set of experiments analyzed those cells bearing both B220 and CD43 for the levels of their expression of HSA and BP-1 (Figures 1 and 2). Flow cytometry plots demonstrated that the B220CD43 cells neatly resolved into three discrete subpopulations. The first, labeled A in Figure 1, expressed neither HSA, nor BP-1. The second, labeled B, expressed HSA, but not BP-1, and the third expressed both of these antigens. Analysis of Ig gene rearrangements in these populations revealed that no gene rearrangements occurred in fraction A but that D to JH gene segment rearrangements had begun in fraction B. Subsequent work has shown that VH to DJH Gated B220+ CD43– FIGURE 2 Flow cytometric characterization of the early developmental stages of B cells. (See text for details.) [Hardy et al., J. Exp. Med. 173, 1213–1225. May 1991. By permission of Rockefeller University Press] (continued) 342 PA R T I V | CLASSIC EXPERIMENT Adaptive Immunity: Development (continued) Start with bone marrow cells Sort for B220+, CD43- mIgMlo H chain rearranged L chain rearrangement beginning mIgMhi H chain rearranged L chains rearranged mIgMhi, mIgDhi H chain rearranged L chains rearranged Fraction D Small pre-B cells Fraction E Immature B cells Fraction F Mature B cells FIGURE 3 The isolation of Hardy’s fractions D, E, and F. Bone marrow cells were sorted for cells bearing B220, but not CD43 (recognized by monoclonal antibody S7) and then analyzed for cell-surface expression of IgM and IgD. Pro-B Cells In the early pro-B-cell stage, D to JH recombination is completed and the cell begins to prepare for V to DJH joining. However, this final recombination event awaits the expression of the quintessential B-cell transcription factor, PAX5. The Pax5 gene is among EBF-1’s transcriptional targets (see Figure 10-7) and transcription of genes controlled by the PAX5 transcription factor denotes passage to the pro-B-cell stage of development, at which point the expression of non-Blineage genes is permanently blocked. PAX5 can act as a transcriptional repressor, as well as an activator, and blocks Notch-1 gene expression, thereby terminating any residual potential of the pro-B cell to develop along the T-cell lineage. Many important B-cell genes are turned on at this stage, under the control of PAX5 and other transcription factors. Among these is the gene encoding CD19, which we first encountered in rearrangements occur in fraction C (although at the time, the method of analysis that Hardy and colleagues used failed to reveal this second type of rearrangement). So at this point, we know that three distinct types of B-cell precursors express both B220 and CD43, and can be discriminated on the basis of their levels of the further two antigens, HSA and BP-1. Fraction A corresponds to what we now know as pre-pro B cells, fraction B to early pro-B cells, and fraction C to late pro-B cells. Culture of fraction C cells yielded cells that expressed membrane (m) IgM; similarly, culture of fraction B cells also yielded daughter cells expressing mIgM, but at a lower frequency than fraction C cells, suggesting that cells in fraction C were further along the differentiation pathway to mIgM B cells. Furthermore, the three different fractions displayed differential dependence on the need to adhere to the stromal cell layer. Cells from fraction A required stromal cell contact for survival. Fraction B cells survived best in contact with the stromal cells, but were able to survive in a culture in which they were separated from the stromal cells by a semipermeable membrane. Under these conditions, they could still receive soluble factors generated by the stromal cells, but were prevented from generating adhesive interactions with stromal cell-surfacebound growth factors. Fraction C cells Chapter 3, as one of the components of the B-cell co-receptor. CD19 is considered a quintessential B-cell marker and is often used as such in flow cytometry experiments. Once the PAX5 protein is expressed, a mutual reinforcement occurs between PAX5 and EBF-1 expression, as illustrated in Figure 10-7, with each transcription factor serving to enhance the expression of the other. The PAX5 protein continues to be expressed in mature B cells until the B cell commits to a plasma-cell fate following antigenic stimulation (see Chapter 12). The higher levels of EBF1 expression induced by PAX5 also allow for an increase in the level of IL-7R expression. PAX5 promotes VH to D recombination by contracting the IgH locus, thus bringing the distant VH gene segments closer to the D-JH region. B cells deficient in PAX5 permit D to JH Ig gene rearrangement, but do not allow recombination of B-Cell Development | CHAPTER 10 343 BOX 10-3 survived and proliferated in the absence of stromal cell contact. Analysis of the factors secreted by the stromal cells that were necessary for survival and proliferation of the fraction B and C cells revealed one of them to be interleukin 7. Hence, using the criteria of Ig gene rearrangements and phenotypic analysis of cultured cell populations, Hardy and colleagues were able to place the three fractions in sequence; cells of fraction A gave rise to cells of fraction B, which in turn mature into cells of fraction C. Careful analysis of the contour graph of fraction C reveals that it, in turn, can be subdivided on the basis of the levels of expression of HSA. That population of cells bearing higher levels of HSA, as well as BP-1, is now defined as fraction C, which corresponds to early or large pre-B cells. Hardy and colleagues next turned their attention to those cells that expressed B220, but had lost CD43, and measured their cell-surface expression of mIgM (Figure 3). Three populations of cells were again evident, which they labeled D, E, and F (Figure 4). Cells belonging to fraction D expressed zero to low levels of mIgM, showed complete heavychain rearrangement, and some lightchain rearrangement and correspond to small pre-B cells. Cells belonging to fraction E displayed high levels of mIgM as well as of B220, complete heavy-chain rearrangement, and most of the cells in that fraction also displayed light-chain gene rearrangement. Fraction E cells are thus immature B cells ready to leave the bone marrow. Subsequent further characterization of fraction F cells showed that, in addition to surface IgM, these cells also bear surface IgD and therefore represent fully mature B cells, presumably recirculating through the bone marrow. Thus, Hardy’s experiments revealed that the pool of progenitor and precursor B cells in the bone marrow represents a complex mixture of cells at different stages of development, with varying requirements for stromal cell contact and interleukin support. These elegant experiments still had one more story to tell that did not appear in the original paper, but which emerged in later publications. Single-cell PCR analysis of fraction C cells showed that many of them had nonproductive rearrangements on both heavy-chain chromosomes (see Figure 2). In contrast, all the cells from the C fraction demonstrated productive rearrangements on one of the heavy-chain chromosomes. Fraction C cells therefore represent B cells that have been unable to productively rearrange one of their heavychain genes and that will therefore eventually die by apoptosis. The C fraction also included the highest proportion of cells in cycle of any of the B220 B-cell stages in the VH to the D Ig gene segment, indicating that PAX5 is essential to the second step of Ig gene rearrangement. Expression of the signaling components of the B-cell receptor, Ig and Ig, also begins at the pro-B-cell stage and the Ig,Ig signaling complex is briefly placed on the cell surface in complex with the chaperone protein calnexin. Although this Ig,Ig complex has been referred to as a “pro-BCR,” no ligand has yet been established for it, nor do we yet understand the importance of any signaling that may emanate from it. Also during the pro-B-cell stage, c-Kit is once more turned on briefly, enabling the cell to receive signals from stem cell factor. By the beginning of the pre-B-cell stage of development, expression of c-Kit is irreversibly turned back off. By the late pro-B-cell stage, most cells have initiated VH to DJH Ig gene segment recombination, which is completed by the onset of the early pre-B-cell stage. Gated B220+ CD43– FIGURE 4 Flow cytometric characterization of the later developmental stages of B cells. (See text for details.) [Hardy et al., J. Exp. Med. 173, 1213–1225. May 1991. By permission of Rockefeller University Press] the marrow. This is consistent with the notion that the new heavy chain is associating with the surrogate light chain at the C stage and the pre-B-cell receptor complex is expressed on the cell surface, triggering the period of clonal expansion of B cells described in this chapter. REFERENCES Hardy, R. R., et al. 1991. Resolution and characterization of pro-B and pre-pro-B cell stages in normal mouse bone marrow. Journal of Experimental Medicine 173:1213–1225. Hardy, R. R., P. W. Kincade, and K. Dorshkind. 2007. The protean nature of cells in the B lymphocyte lineage. Immunity 26:703–714. Pre-B Cells During the pre-B-cell stage, the cell expresses a pre-B-cell receptor composed of the rearranged heavy chain, complexed with the VpreB and 5 components of the surrogate light chain (see Figure 7-10). The appearance of this pre-Bcell receptor signals the entry of the developing B cell into the large, or early pre-B-cell phase. As we learned in Chapter 7, the expression of the heavy chain at the cell surface is necessary for the termination of further heavy-chain rearrangement and ensures allelic exclusion of the Ig heavychain genes. Animals deficient in the expression of either the pre-B-cell receptor, or of the signaling components Ig,Ig, fail to progress to the pre-B-cell stage. Signaling through the pre-B-cell receptor induces a few rounds of proliferation in the pre-B cell. This proliferative 344 PA R T I V | Adaptive Immunity: Development phase correlates in time with the expression on the pre-Bcell surface of CD25, the  chain of the high-affinity IL-2 receptor  chain (see Figure 4-8), which first appears on B cells at the pro-B-cell stage. Since the pre-B-cell proliferative process appears mainly to be driven by IL-7, the functional significance of CD25 at this point is unclear. However, its appearance is frequently used as a marker of the late pro-Bcell to early pre-B-cell stage of development. VH gene recombination is energetically expensive to the organism, as not all developing B cells are successful in undergoing productive VH gene rearrangements, and those that fail to do so are lost by apoptosis. It therefore seems logical that those B cells that have achieved productive heavychain expression should be allowed to proliferate. Each individual daughter cell derived from this proliferative process is then free to participate in a different light-chain rearrangement event. Most individuals will therefore have multiple B cells expressing precisely the same heavy-chain rearrangement but each with a different light chain, and many different receptor specificities can thereby be generated from each successful heavy-chain rearrangement. Recall from Chapter 9 that an analogous process of  chain rearrangement, followed by proliferation prior to  rearrangement, occurs in T cells. If the pre-B-cell receptor cannot be displayed on the cell surface because of non-productive VHDJH gene rearrangements, B-cell development is halted and the cell is lost to apoptosis. This stage in B-cell development is therefore referred to as the pre-B-cell (1st) checkpoint (see Figure 10-6). Progress through this checkpoint depends on some type of signaling event through the pre-B-cell receptor, and recent evidence suggests that this is mediated via interactions between arginine-rich regions in the non-immunoglobulin portion of the 5 component of the surrogate light chain, or in the CDR3 regions of some heavy chains, with negatively charged molecules on the surface of the stromal cells. Pre-B-cell receptor signaling induces the transient downregulation of RAG1/2 and the loss of TdT activity. Together, these events ensure that, as soon as one heavy-chain gene has successfully rearranged, no further heavy-chain recombination is possible. This results in the phenomenon of allelic exclusion, whereby the genes of only one of the two heavychain alleles can be expressed in a single B cell. As a result of this pre-B-cell receptor signaling, the chromatin at the unused heavy-chain locus undergoes a number of physical changes that render it incapable of participating in further rearrangement events. Recall that IL-7 provided one of the signals that brought the VH, D, and JH loci into close apposition with one another at the beginning of VHDJH recombination. A reduction in IL-7 signaling at the pre-B-cell stage now reverses that initial locus contraction, resulting in the physical separation of the VH, D, and JH gene segments in the unrearranged heavychain locus. This decontraction is then followed by deacetylation events that deactivate the unused heavy-chain locus and return it to a heterochromatic (inactive, closed) configuration. Surrogate light-chain expression is also terminated by a negative feedback round of signaling through the pre-B-cell receptor. At the end of pre-BCR-signaled cell proliferation, the pre-B-cell receptor is lost from the surface, and this signals entry into the small or late pre-B-cell stage. At this point, light-chain rearrangement is initiated with the re-expression of the Rag1/2 genes. Very little TdT activity remains at this stage, and therefore N region addition occurs less frequently in light chains than in heavy chains. In the mouse, light-chain rearrangement begins on one of the chain chromosomes, followed by the other. If neither chain rearrangement is successful, rearrangement is then successively attempted on each of the  chain chromosomes. In humans, rearrangement is initiated randomly at either the or the  loci. Once a light-chain gene rearrangement has been successfully completed, the IgM receptor is expressed on the cell surface, signaling entry into the immature B-cell stage. If the attempts at light-chain immunoglobulin gene rearrangement are not successful, the nascent cell is eventually lost at the immature B-cell (2nd) checkpoint (see Figure 10-6). However, given the availability of four separate chromosomes on which to attempt rearrangement, and the opportunity for light-chain editing in the case of unproductive rearrangement, most pre-B cells that have successfully rearranged their heavy chains will progress to the formation of an immature B cell. Immature B Cells in the Bone Marrow Are Exquisitely Sensitive to Tolerance Induction Immature B cells bear a functional receptor in the form of membrane IgM, but have not yet begun to express any other class of immunoglobulin. They continue to express B220, CD25, IL-7R, and CD19. Once the functional BCR is assembled on the B-cell membrane, the receptor must be tested for its ability to bind to self antigens, in order to ensure that as few as possible autoreactive B cells emerge from the bone marrow. Those immature B cells that are found to bear autoreactive receptors undergo one of three fates; some are lost from the repertoire prior to leaving the bone marrow, by the BCR-mediated apoptotic process of clonal deletion. The loss of B cells bearing self-reactive receptors within the bone marrow is referred to as central tolerance. Other autoreactive B cells reactivate their RAG genes to initiate the process of light-chain receptor editing (see Chapter 7). Some autoreactive B cells that recognize soluble self antigens within the bone marrow may survive to escape the bone marrow environment, but become anergic, or unresponsive, to any further antigenic stimuli. The concept of negative selection of lymphocytes bearing autoreactive receptors should be familiar from the discussion of T-cell tolerance in Chapter 9. However, functional differences between T cells and B cells mean that the selection processes against autoreactive B cells are different from those that protect against the emergence of autoreactive T cells, and indeed can be somewhat less stringent. Since stimulation of B-2 B cells requires T-cell help, an autoreactive B cell cannot respond to antigen with antibody production unless there is also an autoreactive T cell that can provide the necessary B-Cell Development cytokines and costimulation (Chapter 12). Thus, it is quite possible that most individuals carry significant numbers of autoreactive B cells within their mature B-cell repertoires that are never activated. There are also mechanistic differences between the modes of negative selection among B and T cells. At this point, no equivalent of the AIRE protein has been shown to exist for B-cell selection, and so B-cell negative selection within the bone marrow is more limited with respect to the available tolerogenic antigenic specificities than is T-cell negative selection in the thymus. Many, but Not All, Self-Reactive B Cells Are Deleted within the Bone Marrow Our understanding of how the immune system eliminates or neutralizes autoreactive threats has been facilitated by the development of transgenic animals which express both deliberately introduced auto-antigens and the receptors that recognize them. It has long been established that crosslinking the IgM receptors of immature B cells in vitro (performed experimentally by treating the cells with antibodies against the receptor  chain) results in death by apoptosis. In contrast, performing the same experiments with mature B cells, bearing both IgM and IgD receptors, results in activation. David Nemazee and colleagues set out to test whether the apoptotic response of immature B cells in vitro reflected what happens in the bone marrow in vivo when an immature B cell meets a self antigen. Nemazee et al.’s approach was conceptually simple, although experimentally complex, particularly for the time period in which the work was done (1989). They generated mice transgenic for both a heavy and a light chain specific for the MHC molecule H-2Kk. All the B cells in this mouse therefore made only anti-H-2Kk antibodies. If immature B cells undergo selection to prevent autoimmunity, these cells would be selected against in a mouse that expresses the H-2Kk gene for MHC. By appropriate breeding, they introduced the immunoglobulin H-2Kk-specific transgenes into mice bearing two different MHC genotypes. In the first group of mice (Figure 10-8a), which bore H-2Kd but no H-2Kk antigens, they were able to detect the transgenic antibody at high frequency on the surface of B cells and at high concentration in the serum (Table 10-1). This makes sense, as the transgenic antibody would be unable to bind the H-2Kd molecules and so the B cells that produce it would not be negatively selected. However, when they bred these animals with mice of the H-2Kk type (Figure 10-8b), no membrane-bound or secreted anti-H-2Kk antibodies could be detected, suggesting that all immature B cells bearing the potentially autoimmune receptor antibodies had been deleted in the bone marrow. This deletion occurred via induction of apoptosis in the autoimmune cells. Interestingly, in the H-2Kk/d mice, not all B cells bearing the autoimmune transgenes were deleted, even though all B cells in this mouse should bear the anti-H-2Kk receptor | CHAPTER 10 345 (Figure 10-8c). Closer examination revealed that some of the residual transgene-expressing B cells in the bone marrow had undergone light-chain receptor editing (see Chapter 7), changing their antigen specificity so they no longer bound the H-2Kk antigen. Recent experiments suggest that in vivo, a significant fraction of potentially autoimmune B cells undergo receptor editing (or even VH gene replacement (Chapter 7), and successfully generate acceptable BCRs prior to release from the bone marrow. In normal animals, not all potentially autoimmune B cells are lost to clonal deletion or altered via receptor editing or VH gene replacement within the bone marrow, however; some are released to the periphery and subject to further rounds of selection. B Cells Exported from the Bone Marrow Are Still Functionally Immature Once the B cell expresses IgM on its membrane (mIgM), it is referred to as an immature B cell. This B cell is ready for export to the spleen, where it completes its developmental program. Immature B cells have a short half-life, in part as a result of expressing low levels of the anti-apoptotic molecules Bcl-2 and Bcl-xl. They also express high levels of the cell-surface molecule, Fas, which is capable of transmitting a death signal when bound by its ligand (see Chapters 4 and 9). Immature B cells are exquisitely susceptible to tolerance induction, and if they encounter a self antigen at this stage of development, the B cells will re-express the RAG1 and RAG2 genes and edit their light-chain genes. If receptor editing fails to yield a suitable receptor, the cell undergoes apoptosis. The study of B-cell development in the periphery, like that in the bone marrow, has benefited significantly from the ingenious application of flow cytometry, which has enabled the classification of immature B cells into two subpopulations of transitional B cells (T1, T2). These transitional B cells act sequentially as the precursors to the fully mature B cell. T1 and T2 Transitional B Cells T1 and T2 transitional B cells were characterized initially on the basis of their cell-surface expression of immunoglobulin receptors and membrane markers (Table 10-2). T1 cells are mIgMhi, mIgD/lo, CD21, CD23, CD24, and CD93. T2 cells differ from T1 cells in having higher levels of mIgD and in expressing CD21 (the complement receptor and B-cell co-receptor; see Figure 3-7) and CD23. T2 cells also express BAFF-R, the receptor for the B-cell survival factor BAFF, whose expression is dependent on signals received through the BCR. As B cells differentiate from the transitional T2 state to full maturity, they raise their levels of mIgD still further, while reducing the expression of mIgM. They also cease to express CD24 and CD93. T1 cells that have been labeled and transferred into recipient mice develop into T2 cells. Similarly, both transitional B-cell subpopulations have been demonstrated to have the 346 PA R T I V | Adaptive Immunity: Development (a) H-2d transgenics Mature B cells express anti -Kk Kd (b) H-2d/k transgenics Immature B cells Anti-Kk Kd No mature B cells express anti-Kk Kk Bone marrow stromal cell (c) H-2d/k transgenics Light-chain editing A few mature B cells with new light chains no longer bind Kk FIGURE 10-8 Experimental evidence for negative selection (clonal deletion) of self-reactive B cells during maturation in the bone marrow. The presence or absence of mature peripheral B cells expressing a transgene-encoded IgM against the H-2 class I molecule Kk was determined in H-2d mice (a) and H-2d/k mice (b) and (c). (a) In the H-2d transgenics, the immature B cells did not bind to a self antigen and consequently went on to mature, so that splenic B cells expressed the transgene-encoded anti-Kk as membrane Ig. (b) In the TABLE 10-1 H-2d/k transgenics, many of the immature B cells that recognized the self antigen Kk were deleted by negative selection. (c) More detailed analysis of the H-2d/k transgenics revealed a few peripheral cells that expressed the transgene-encoded  chain but a different light chain. Apparently, a few immature B cells underwent light-chain editing, so they no longer bound the Kk molecule and consequently escaped negative selection. [Adapted from D. A. Nemazee and K. Burki, 1989, Nature 337:562; S. L. Tiegs et al., 1993, Journal of Experimental Medicine 177:1009.] Expression of transgene encoding IgM antibody to H-2k class I MHC molecules Expression of transgene Experimental animal Number of animals tested As membrance Ab As secreted Ab (g/ml) Nontransgenics 13 () 0.3 H-2d transgenics 7 () 93.0 H-2d/k transgenics 6 () 0.3 [Adapted from D. A. Nemazee and K. Bürki, 1989, February, Clonal deletion of B lymphocytes in a transgenic mouse bearing anti-MHC class I antibody genes, Nature 337:562–566. doi:10.1038/337562a0.] B-Cell Development | CHAPTER 10 347 Bone marrow sinusoids, bloodstream Endosternum T cell zone Pro and pre B Immature B Follicle Central arteriole Bone marrow T2 IgM IgD T1 Marginal sinus Marginal zone FIGURE 10-9 T2, but not T1, transitional immature cells can enter the B-cell follicles and recirculate. Immature B cells leave the bone marrow as T1 transitional immature B cells. They enter the spleen from the bloodstream through the marginal sinuses, percolating into the T-cell zones, and differentiating into T2 transitional B cells, which gain the ability to enter the B-cell follicles and recirculate. There, the T2 cells complete their differentiation into mature follicular B-2 cells. Marginal zone cells have also been shown to derive from T2 cells. [Adapted from J. B. Chung, M. Silverman, and J. G. Monroe, 2003, Transi- capacity to differentiate into mature B cells. These experiments have therefore proven that the order of the developmental sequence progresses from T1 to T2 to mature B cell. The time in transit of a T1 cell to a mature B cell has been measured to be approximately 3 to 4 days. Most T1 B cells differentiate to T2 cells within the spleen, but a minority of about 25% of transitional B cells emerge from the bone marrow already in the T2 state. The increased level of maturity of T2 cells correlates with changes in the expression of chemokine and cytokine receptors, such that T2 cells, but not T1 cells, are capable of recirculating among the blood, lymph nodes, and spleen; and T2 cells, but not T1 cells, can enter B-cell follicles. Figure 10-9 shows the path of the developing transitional B cell as it leaves the bone marrow and enters the spleen through the central arteriole, which deposits it in the marginal sinuses, just inside the outer marginal zone. (In humans, the anatomy of the spleen is slightly different, and the cells arrive in the spleen in a peri-follicular zone.) From there, the T1 B cell percolates through to the T-cell zone, where some fraction of T1 cells will mature into the T2 state. T2 B cells are then able to enter the follicles or the marginal zone where they complete their developmental program into fully mature, recirculating B lymphocytes. In Chapter 7, we learned that mature B cells bear on their surfaces two classes of membrane-bound immunoglobulins— IgM and IgD—and that the expression of mIgD along with mIgM requires carefully regulated mRNA splicing events. It is at the point of transit between the T1 and T2 stages of development that we observe the onset of these splicing capabilities. Mature B cells bear almost 10 times more mIgD than mIgM, and so mIgD expression results in significant up-regulation in the number of B-cell immunoglobulin receptors. The effect of strong BCR engagement with a multivalent, or membrane-bound, antigen depends on the maturational status of the transitional B cell (Figure 10-10 and Table 10-3). Selfreactive T1 B cells are eliminated by apoptosis in response to a strong antigenic signal, in a process reminiscent of thymocyte negative selection, leading to peripheral tolerance; recent experiments have suggested that in healthy adults, fully 55% to 75% of immature B cells are lost by this process. In contrast, once the B cell has matured into a T2 transitional B cell, it becomes resistant to antigen-induced apoptosis, reminiscent of thymocytes that have reached the single-positive stage of development. This TABLE 10-2 Surface marker expression on transitional T1 and T2 and mature B-2 B cells T1 T2 Mature B-2 cells mIgM High High Intermediate mIgD  Low Intermediate High CD24    CD93    CD21    CD23    /   Marker BAFF receptor (BAFF-R) Note: CD93 is defined by the AA4 monoclonal antibody. CD24 is otherwise known as the Heat Stable Antigen (HSA). CD23 is a low-affinity receptor for IgE. CD21 is a receptor for complement and part of the B-cell co-receptor. After D. Allman and S. Pillai. 2008. Peripheral B cell subsets. Current Opinion in Immunology 20:149–157, and others. tional B cells: Stem by step towards immune competence, Trends in Immunology 24:343–348, Figure 1; and R. E. Mebius, and G. Kraal, 2005, Structure and function of the spleen. Nature Reviews Immunology 5:606–616, Figure 1.] PA R T I V 348 PALS | Adaptive Immunity: Development Follicle T1 T2 Negative selection Positive selection Death Death B-2 FIGURE 10-10 Transitional B cells bound for a follicular fate undergo positive and negative selection in the spleen. T1 transitional B cells, which recognize antigen with high affinity in the spleen, are eliminated by negative selection, and never reach the splenic follicles. Those T1 cells that escape negative selection enter the follicles and differentiate into T2 B cells. In the follicles, their BCRs interact with an unknown molecule(s) that deliver(s) a stimulatory survival signal. Transitional cells that have received this survival signal up-regulate their BAFF receptors (positive selection). Those T2 B cells that fail to receive the stimulatory signal (or that fail to receive a BAFF survival signal) die in the spleen. Selecting antigens are shown as violet shapes; T1 and T2 cells are green. White cells represent dead cells that were either negatively selected or failed positive selection. [Adapted from T. T. Su et al., 2004, Signaling in transitional type 2 B cells is critical for peripheral B-cell development, Immunological Reviews 197:161–178, Figure 3.] resistance to receptor-induced cell death results in part from the fact that T2 B cells have increased their expression of the anti-apoptotic molecule Bcl-xl. Using conditional knockout genetic techniques (see Chapter 20), animals can be generated that lack Ig receptor TABLE 10-3 Responses to strong BCR signaling in T1, T2, and mature B-2 B cells Nature of Response T1 T2 Mature B-2 B cells Formation of lipid rafts /   Increase in cytoplasmic Ca2 ion concentrations    Increase in diacylglycerol concentrations    Induction of Bcl-xl    Induction of apoptosis    expression at different stages of B-cell development. Animals that fail to express a BCR during the immature B-cell stage lose the capacity to make any mature B cells at all. This indicates that some low level of tonic signaling through the BCR, analogous to the positive selection signal needed by developing thymocytes, is required for continued generation and survival of immature B cells. B cells unable to receive this signal die at the T2 stage (see Figure 10-10). Those T2 B cells able to receive the follicular signal up-regulate the expression of the receptor for the B-cell survival factor BAFF. Given the different outcomes of signaling through the BCR for T1 versus T2 B cells, it is clear that there must be differences in the signaling pathways downstream of the BCR in the two transitional B-cell types. Specifically, BCR-mediated signaling of T1 B cells results in calcium release without significant production of diacylglycerol, and provides an apoptotic signal. In contrast, receipt of BCR signals by T2 B cells induces both an increase in the concentration of intracytoplasmic calcium and in diacylglycerol production. This combination of intracellular second messengers delivers both maturational and survival signals to the cell and suggests the involvement of a diacylglycerol-activated protein kinase in survival signaling (see Chapter 3). But what causes this difference in the signal transduction pathways between T1 and T2 cells? A partial answer to this question appears to lie in differences in the composition of the lipid membranes of the two types of cells. Immature T1 B cells contain approximately half as much cholesterol as their more mature counterparts, and this reduction in cholesterol levels appears to prevent efficient clustering of the B-cell receptor into lipid rafts upon BCR stimulation. This may cause a reduction of the strength of BCR signaling in T1 versus T2 immature B cells. The development of B cells through the transitional phase is absolutely dependent on signaling through the BAFF receptor (BAFF-R). BAFF-R expression is first detected in T1 B cells and increases steadily thereafter. BAFF is then required constitutively throughout the life of mature B cells. Signaling through the BAFF/BAFF-R axis promotes survival of transitional B cells by inducing the synthesis of anti-apoptotic factors such as Bcl2, Bcl-xl, and Mcl-1, as well as by interfering with the function of the pro-apoptotic molecule Bim (see Figure 9-12). The discovery of a B-cell survival signal mediated by BAFF/BAFF-R interactions, and distinct from survival signals emanating from the BCR, extends our thinking about how B cells are selected for survival in the periphery. In the presence of high levels of BAFF, B cells that may not otherwise receive sufficient quantities of survival signals via the BCR may survive a selection process that would otherwise eliminate them. In this way, BAFF can provide plasticity and flexibility in the process of B-cell deletion. However, this may be accomplished at the cost of maintaining potentially autoreactive cells. T3 B Cells Are Primarily Self-Reactive and Anergic Transitional T3 B cells were first characterized in the blood and lymphoid organs by flow cytometry and were described B-Cell Development as being CD93mIgMlowCD23. The function of CD93 is so far unknown; CD23 is a low-affinity receptor for some classes of immunoglobulin, and the two markers were used in these experiments only to identify the cell populations, and not because of any particular reasons relevant to their functionality. Recent experiments have suggested that the T3 population may represent B cells that have been rendered anergic by contact with soluble self antigen but have not yet been eliminated from the B-cell repertoire. A transgenic system developed by Goodnow and colleagues first placed the concept of B-cell anergy, or unresponsiveness, onto a firm experimental footing. Anergic lymphocytes clearly recognize their antigens, as shown by the identification of low levels of molecular signals generated within the cells after binding to antigen. However, rather than being activated by antigen contact, anergic B cells fail to divide, differentiate, or secrete antibody after stimulation, and many die a short time after receipt of the antigenic signal. Goodnow et al. developed the two groups of transgenic mice illustrated in Figure 10-11a. One group of mice carried a hen egg-white lysozyme (HEL) transgene linked to a metallothionine promoter, which placed transcription of the HEL gene under the control of zinc levels in the animals’ diet. This allowed the investigators to alter the levels of soluble HEL expressed in the experimental animals by changing the concentration of zinc in their food. Under these experimental conditions, HEL was expressed in the periphery of the animal, but not in the bone marrow. The other group of transgenic mice carried rearranged immunoglobulin heavy-chain and light-chain transgenes encoding anti-HEL antibody; in these transgenic mice, the rearranged anti-HEL transgene is expressed by 60% to 90% of the mature peripheral B cells. Goodnow then mated the two groups of transgenics to produce “double-transgenic” offspring carrying both the HEL and anti-HEL transgenes (Figure 10-11b) and asked what effect peripheral HEL expression would have on antibody expression by B cells bearing the anti-HEL antibodies. They found that the double-transgenic mice continued to generate mature, peripheral B cells bearing anti-HEL membrane immunoglobulin of both the IgM and IgD classes, indicating that the B cells had fully matured. However, these B cells were functionally nonresponsive, or anergic. Flow-cytometric analysis of B cells from the doubletransgenic mice showed that, although large numbers of anergic anti-HEL cells were present, they expressed membrane IgM at levels about 20-fold lower than anti-HEL single transgenics (Figure 10-11b). When these mice were given an immunizing dose of HEL, few anti-HEL plasma cells were induced and the serum anti-HEL titer was very low (Table 10-4). Furthermore, when antigen was presented to these anergic B cells in the presence of T-cell help, many of the anergic B cells responded by undergoing apoptosis. Additional analysis of the anergic B cells demonstrated that they had a shorter halflife than normal B cells and appeared to be excluded from the B-cell follicles in the lymph nodes and spleen. These properties were dependent on the continuing presence of | CHAPTER 10 349 antigen, as the B cell half-lives were restored to normal lengths upon adoptive transfer of the transgene-bearing B cells to an animal that was not expressing HEL. The anergic response appears to be generated in vivo when immature B cells meet a soluble self antigen. More recent experiments have focused on defining the differences between the signal transduction events leading to anergy versus activation. Anergic B cells show much less antigen-induced tyrosine phosphorylation of signaling molecules, when compared with their nonanergic counterparts and antigen-stimulated calcium release from storage vesicles into the cytoplasm of the anergic B cells was also dramatically reduced. Anergic B cells also require higher levels of the cytokine BAFF for continued survival, and it is likely that their reduced halflives result from unsuccessful competition with normal B cells for limiting amounts of this survival molecule. One of the outcomes of BAFF signaling is a reduction in the cytoplasmic levels of the pro-apoptotic molecule BIM; as might be expected, anergic B cells show higher-than-normal levels of BIM and a correspondingly increased susceptibility to apoptosis. The conclusion from these experiments is that mechanisms exist, even after the B cells have exited the bone marrow and entered the periphery, that minimize the risk that B cells make antibodies to soluble self proteins expressed outside the bone marrow. B cells reactive to such proteins respond to receptor stimulation in the absence of appropriate T-cell help by anergy and eventual apoptosis. What might be the function of these anergic cells? One possibility is that they serve to absorb excess self antigens that might otherwise be able to deliver activating signals to highaffinity B cells and thus lead to autoimmune reactions. Another is that they represent cells destined for apoptosis that do not yet display the characteristic microanatomy of apoptotic cells. Yet another is that these cells will eventually develop into B regulatory cells (see Chapter 12). As is so often the case in the immune system, it is more than possible that all of these functions are subsumed within this intriguing cell population, which remains the subject of intensive current investigation. Mature, Primary B-2 B Cells Migrate to the Lymphoid Follicles Fully mature B cells express high levels of IgD and intermediate levels of IgM on their cell surfaces (see Table 10-2). Mature B cells recirculate between the blood and the lymphoid organs, entering the B-cell follicles in the lymph nodes and spleen, and responding to antigen encounter in the presence of T-cell help with antibody production (Chapter 12). Approximately 10 million to 20 million B cells are produced in the bone marrow of the mouse each day, but only about 10% of this number ever take up residence in the periphery and only 1% to 3% will ever enter the recirculating follicular B-2 B-cell pool. Some of these cells are lost to the process of clonal deletion, but others are perfectly harmless B cells that nonetheless fail to thrive. Experimental depletion of the mature B-cell population, either chemically or by irradiation, 350 PA R T I V | Adaptive Immunity: Development (a) × Transgenic (HEL) Transgenic (anti–HEL) Double transgenic (carrying both HEL and anti–HEL transgenes) HEL-binding B cells (b) Nontransgenic Anti–HEL transgenic Anti–HEL/HEL double transgenic 100 10 1 1 1 10 100 1 10 10 100 IgM expression on membrane (arbitrary fluorescence units) 100 onstrating clonal anergy in mature peripheral B cells. (a) Production of double-transgenic mice carrying transgenes encoding HEL (hen egg-white lysozyme) and anti-HEL antibody. (b) Flowcytometric analysis of peripheral B cells that bind HEL compared with membrane IgM levels. The number of B cells binding HEL was measured by determining how many cells bound fluorescently labeled HEL. Levels of membrane IgM were determined by incubating the cells with anti-mouse IgM antibody labeled with a fluorescent label different from that used to label HEL. Measurement of the fluorescence emitted from this label indicated the level of membrane IgM expressed by the B cells. The nontransgenics (left) had many B cells that expressed high levels of surface IgM but almost no B cells that bound HEL above the background level of 1. Both antiHEL transgenics (middle) and anti-HEL/HEL double transgenics (right) had large numbers of B cells that bound HEL (blue), although the level of membrane IgM was about 20-fold lower in the double transgenics. The data in Table 10-4 indicate that the B cells expressing anti-HEL in the double transgenics cannot mount a humoral response to HEL. followed by in vivo reconstitution, results in rapid replenishment of the B-cell follicular pool. This suggests that the follicular B-cell niches have a designated capacity and that once full they turn away additional B cells. Most probably, the mechanism for this homeostatic control of B-cell numbers relies on competition for survival factors, particularly BAFF and its related proteins. Experiments using conditional RAG2 knockout animals, in which all new B-cell development was prevented in otherwise healthy adult animals, indicate that follicular B-2 B cells have a half-life of approximately 4.5 months. In contrast, since B-1 B cells can self-renew in the periphery, their numbers are unaffected in this experimental knockout animal. FIGURE 10-11 Goodnow’s experimental system for dem- TABLE 10-4 Expression of anti-HEL transgene by mature peripheral B cells in single- and double-transgenic mice Experimental Group Anti-HEL single transgenics Anti-HEL/HEL double transgenics HEL level Membrane anti-HEL Ig Anti-HEL PFC/spleen* Anti-HEL serum titer None  High High 9  Low Low 10 M * Experimental animals were immunized with hen egg-white lysozyme (HEL). Several days later hemolytic plaque assays for the number of plasma cells secreting antiHEL antibody were performed and the serum anti-HEL titers were determined. PFC plaque-forming cells. Adapted from Goodnow, C. C., 1992, Annual Review of Immunology 10:489. | B-Cell Development The Development of B-1 and Marginal-Zone B Cells This chapter has so far focused on the development of those B cells that belong to the best characterized B-cell subpopulation, B-2 B cells (or follicular B cells). Mature B-2 B cells recirculate between the blood and the lymphoid organs, and can be found in large numbers in the B-cell follicles of the lymph nodes and spleen. However, other subsets of B cells have been recognized that perform distinct functions, occupy distinct anatomical locations, and pursue different developmental programs. This section of the chapter will therefore address the development and function of B-1 B cells and of marginal-zone B cells (see Figure 10-12 for a comparison of the properties of these three cell types). As described more fully in Chapter 12, B-1 B cells generate antibodies against antigens shared by many bacterial species and may do so even in the absence of antigenic stimulation. They are the source of the so-called natural antibodies: serum IgM antibodies that provide a first line of protection against invasion by many types of microorganisms. Marginal-zone B cells take their name from their location in the outer zones of the white pulp of the spleen. They are the first B cells encountered by blood-borne antigens 351 entering the spleen and, like B-1 B cells, mainly (although not exclusively) produce broadly cross-reactive antibodies of the IgM class. Both B-1 and marginal-zone B cells can generate antibodies in the absence of T-cell help, although the addition of helper T cells enhances antibody secretion and allows for some degree of heavy-chain class switching. B-1 B Cells Are Derived from a Separate Developmental Lineage B-1 B cells are phenotypically and functionally distinct from B-2 B cells in a number of important ways. They occupy different anatomical niches from B-2 B cells, constituting 30% to 50% of the B cells in the pleural and peritoneal cavities of mice, and representing about 1 million cells in each space. A similar number of B-1 B cells can also be found in the spleen, but there they represent a much smaller fraction (around 2%) of the splenic B-cell population. B-1 B cells have only a relatively limited receptor repertoire, and their receptors tend to be directed toward the recognition of commonly expressed microbial carbohydrate antigens. These broadly cross-reactive, low-affinity antigen receptors expressing minimal repertoire diversity are reminiscent of the pathogen-associated molecular pattern (PAMP) receptors of the innate immune system Marginal zone B cells B-1 B cells Follicular (B-2) B cells CD19/ CD21 CHAPTER 10 IgM CD19/ CD21 IgM IgM IgD CD1 CD5 CD23 (B-1a cells only) Attribute Follicular (B-2) B cells B-1 B cells Marginal zone B cells Major sites Secondary lymphoid organs Peritoneal and pleural cavities Marginal zones of spleen Source of new B cells From precursors in bone marrow Self-renewing (division of existing B-1 cells) Long-lived May be self-renewing V-region diversity Highly diverse Restricted diversity Somewhat restricted Somatic hypermutation Yes No Unclear Requirements for T-cell help Yes No Variable Isotypes produced High levels of IgG High levels of IgM Primarily IgM; some IgG Response to carbohydrate antigens Possibly Yes Yes Response to protein antigens Yes Possibly Yes Memory Yes Very little or none Unknown Surface IgD on mature B cells Present on naïve B cells Little or none Little or none FIGURE 10-12 The three major populations of mature B cells in the periphery. The cell-surface properties and functions of B-2, B-1, and marginal-zone (MZ) cells are shown. Conventional B-2 cells were so named because they develop after B-1 B cells. 352 PA R T I V | Adaptive Immunity: Development (Chapter 5), and B-1 B cells are thus considered to play a role that bridges those of the innate and adaptive immune systems. In contrast with the T-1 transitional B-2 B cells, which undergo apoptosis upon antigen challenge, transitional B-1 cells undergo apoptosis unless they interact with self antigens. Work in a number of different transgenic systems has suggested that relatively strong BCR engagement by self antigens provides a positive selection rather than a negative selection survival signal for B-1 B cells. In contrast with B-2 cells and marginalzone cells (see below), B-1 B cells do not require interaction with BAFF during the transitional stage of development. For many years, the preeminent issue debated by those interested in the development of B-1 B cells was whether they constituted a separate developmental lineage, or whether they derived from the same progenitors as B-2 B cells. This controversy has since been resolved in favor of the assertion that B-1 and B-2 B cells derive from distinct lineages of progenitor cells. Several lines of evidence support this conjecture: • B-1 B cells appear before B-2 B cells during ontogenic development. B cells generated from the AGM region and the liver in the fetus have a cell-surface phenotype characteristic of B-1 B cells and secrete natural IgM antibodies without the need for deliberate immunization. B-1 B cells may be generated early in development in order to protect the fetus from commonly encountered bacterial pathogens. • B-1 B cells display much more limited V region diversity than B-2 B cells. Generated at a point in ontogeny before TdT can be efficiently activated, many B-1 cells lack any evidence of N region addition. • B-1 B cells populate different anatomical niches in the mouse than the later-arising B-2 B cells, in much the same manner that has been described for the fetally derived  T cells (Chapter 9). • B-cell progenitors of the CD19CD45Rlow/ phenotype transferred into an immunodeficient mouse were able to repopulate the B-1, but not the B-2 B-cell compartments. Conversely, CD19CD45Rhi B-cell progenitors gave rise to B-2, but not to B-1, daughter cells in an immunodeficient recipient mouse, supporting the notion that the two subclasses derive from different lineages of progenitor cells. • Whereas B-2 B cells must be constantly replenished by the emergence of newly generated cells from the bone marrow, B-1 B cells are constantly regenerated in the periphery of the animal. Bone marrow ablation therefore leaves the mouse with a depleted B-2 pool, but with a fully functional B-1 population. There are very few absolutes in biology, and so it should be clearly stated that it is probable that not all B-1 B cells in an adult animal are derived from fetal liver precursors and that some replenishment of B-1 precursors occurs in the adult bone marrow. Not all B-1 B cells are restricted to IgM production, and the antibody production of some B-1 B cells does benefit from the provision of T-cell help. But notwithstanding these variations on the general themes we have elaborated, it remains the case that evidence clearly supports the notion that B-1 B cells are derived from a different progenitor cell lineage from B-2 B cells. Marginal-Zone Cells Share Phenotypic and Functional Characteristics with B-1 B Cells and Arise at the T2 Stage The term marginal zone refers to the fact that marginal zone (MZ) B cells are located in the outer regions of the white pulp of the spleen (Figure 10-13). The spreading out of the fastmoving arterial flow into the marginal sinuses results in a decrease in the rate of blood flow and allows blood-borne antigens to interact with cells resident within the marginal zone. Indeed, MZ B cells appear to be specialized for recognizing blood-borne antigens. They are capable of responding to both protein and carbohydrate antigens, and evidence suggests that some, but maybe not all, MZ B cells can do so without the need for help from T cells. MZ cells are characterized by relatively high levels of membrane IgM, and the complement receptor/B-cell coreceptor CD21 (see Chapters 3 and 6), but low levels of membrane IgD and the Fc receptor CD23. They also display phospholipid receptors and adhesion molecules that enable them to make adhesive interactions with other cells within the marginal zone that hold them in place. MZ cells are long lived and may self-renew in the periphery. What drives an immature B cell down the MZ pathway? Phenotypic characterization of bone marrow and peripheral immature B-cell populations, combined with labeling studies that define precursor/daughter-cell relationships have suggested that MZ and follicular B-2 B cells both derive from the T2 transitional population (see Figure 10-9). MZ cells, like B-2 B cells, are also reliant on the B-cell survival factor BAFF, which binds to the MZ B cells through the receptor BR3. Like developing B-1 B cells, developing MZ cells appear to require relatively strong signaling through the BCR in order to survive. Surprisingly, unlike any other B-cell subset so far described, the differentiation of MZ B cells also requires signaling through ligands of the Notch pathway (see Chapter 9). Loss of the Notch 2 receptor or deletion of the Notch 2 ligand, Delta-like 1 (Dl-1), results in the selective deletion of MZ B cells. Both MZ and B-1 B cell populations are enriched in cells that express self antigen-specific receptors, and so relatively strong signaling of T2 cells by binding of self antigens through the BCR is also necessary for MZ B-cell differentiation. Comparison of B- and T-Cell Development In closing this chapter, it is instructive to consider the many points of comparison between the development of the two arms of the adaptive immune system, B cells and T cells (Table 10-5). Both cell lineages have their beginnings in the fetus and the neonate. In the neonate,  T cells and B-1 B cells are dispatched to their own particular peripheral niches, Central arteriole T cell zone Follicle B cell zone Marginal sinus Marginal zone FIGURE 10-13 The relative locations of the marginal zone and the follicles in the spleen. This figure shows a cross-section of the spleen, displaying the anatomical relationships between the central arteriole, the T-cell and B-cell zones, and the marginal zone. TABLE 10-5 Comparison between T-cell and B-cell development Structure or process Development begins in the bone marrow. B cells T cells   Development continues in the thymus.   Ig heavy chain or TCR chain many gene rearrangement begins with D-J and continues with V-DJ recombination.   The H chain (BCR) or  chain (TCR) is expressed with a surrogate form of the light chain on the cell surface. Signaling from this pre-BCR or pre-TCR is necessary for development to continue.   Signaling through the pre-B- or pre-TCRs results in proliferation.   The / (BCR) or  (TCR) chain bears only V and J segments.   Signaling from the completed receptor is necessary for survival (positive selection).   / True for the minority B-1 B-cell subset. Ligand for B-2 and MZ subsets unclear.  Low-affinity binding of self MHC and self peptide in the thymus is necessary for positive selection. Receptor editing of / (BCR) or  (TCR) chain modifies autoreactive specificities.  Has been shown to occur, but rarely used as a mechanism of escaping autoreactivity. Immature cells bearing high affinity autoreactive receptors are eliminated by apoptosis (negative selection).   Negative selection involves ectopic expression of auto-antigens in the primary lymphoid organs.  Negative selection involves recognition of MHC-presented peptides.  Positive selection requires recognition of self components.  Auto-antigens are expressed in the thymus under the influence of the AIRE transcription factor.  Heavy or TCR chain allelic exclusion.  Light ( /) or TCR chain allelic exclusion.  90% of lymphocytes are lost prior to export to the periphery.   Development is completed in the periphery to allow for tolerance to antigens not expressed in the primary lymphoid organs    Many T cells display more than one TCR chain. Note: The comparisons in this table most accurately refer to the predominant  T and follicular B-2 cell subsets, although many of the statements made are equally valid for the minority lymphocyte subsets. 353 354 PA R T I V | Adaptive Immunity: Development there to function as self-renewing populations until the death of the host. In the adult animal, B-cell and T-cell development continue in the bone marrow, starting with hematopoietic stem cells. B and T cells share the early phases of their developmental programs, as they pass through progressively more differentiated stages as MPPs, LMPPs, and ELPs. At the ELP and CLP stages, T-cell progenitors leave the bone marrow, and migrate to the thymus to complete their development, leaving B-cell progenitors behind. In mammals, B cells do not have an organ analogous to the thymus in which to develop into mature, functioning cells, although birds do possess such an organ—the bursa of Fabricius. (Many students are unaware that it is from this organ, and not from the bone marrow, that the “B” in B cells originates.) With the initiation of V(D)J recombination, the B cell irreversibly commits to its lineage, and begins the process of receptor rearrangement, followed by B-cell selection and differentiation that will culminate in the formation of a complete repertoire of functioning peripheral B cells. B cells and T cells must both pass through stages of positive selection, in which those cells capable of receiving survival signals are retained at the expense of those which cannot. They must also survive the process of negative selection, in which lymphocytes with high affinity for self antigens are deleted. The process of positive selection in B cells—its mechanism, and the BCR ligands involved— remains one of the least well characterized processes in B-cell development. Unlike T cells, however, B cells do not undergo selection with respect to their ability to bind to self MHC antigens. With the expression of high levels of IgD on the cell surface and the necessary adhesion molecules to direct their recirculation, development of the mature, follicular B-2 cell is complete and, for a few weeks to months, it will recirculate, ready for antigen contact in the context of T-cell help and subsequent differentiation to antibody production. For the final, antigen-stimulated stages of B-cell differentiation, the reader is directed to Chapter 12. S U M M A R Y ■ ■ ■ ■ ■ ■ ■ Hematopoiesis in the embryo generates rapidly dividing hematopoietic stem cells that populate the blood system of the animal, providing red blood cells that supply the oxygen needs of the fetus, and generating the early precursors of other blood-cell lineages. Early sites of blood-cell development include the yolk sac and the placenta, as well as the aorta-gonad-metanephros (AGM) region and the fetal liver. The earliest B-cell progenitors in the fetal liver supply B-1 B cells that migrate into the pleural and peritoneal cavities and remain self-renewing throughout the life of the animal. In adult animals, hematopoiesis occurs in the bone marrow. Progenitor stages of the conventional B-2 subset are defined by the presence of particular cell-surface markers, which include chemokine and lymphokine receptors and proteins involved in adhesive interactions. Progenitor stages are also defined by the status of immunoglobulin V region gene rearrangements. The heavychain V genes rearrange first, with D to JH recombination occurring initially, followed by VH to DJH recombination. The heavy chain is then expressed on the cell surface in combination with the surrogate light chain, which is made up of VpreB and 5. Together they form the pre-B-cell receptor, which is expressed on the cell surface along with the Ig,Ig signaling complex. Signaling through the B-cell receptor stops VH gene rearrangement and calls for a few rounds of cell division. This allows multiple B cells to use the same, successfully rear- ranged heavy chain in combination with many different light chains. ■ ■ ■ ■ ■ After light-chain rearrangement, and the expression of the completed immunoglobulin receptor on the cell surface, immature B cells specific for self antigens present in the bone marrow are deleted by apoptosis. Immature B cells emerge from the bone marrow as transitional 1 (T1) B cells and circulate to the spleen. Interaction with self antigens in the spleen can give rise to apoptosis. T1 B-2 B cells then enter the follicles, where the level of IgD expression increases, they become T2 cells, and then mature into either a follicular B-2 cell (a conventional B cell) or a marginal-zone (MZ) B cell. The three major subsets of B cells differ according to their site of generation, their sites of maturation, their anatomical niches in the adult, the antigens to the which they respond, their need for T-cell help in antibody production, the diversity of their immunoglobulin repertoires, and their abilities to undergo somatic hypermutation and memory generation following antigenic stimulation. Like T cells, developing B cells must undergo both positive and negative selection. Unlike T cells, B cells need not be selected for their ability to recognize antigens in the context of MHC antigens, nor is there a primary immune organ aside from the bone marrow specialized for their maturation. There is no equivalent of the AIRE protein that has yet been discovered to provide for ectopic expression of antigens in the bone marrow in order to facilitate clonal deletion of self antigen-specific B cells. B-Cell Development | CHAPTER 10 355 R E F E R E N C E S Allman, D., and Pillai, S. 2008. Peripheral B cell subsets. Current Opinion in Immunology 20:149–157. Cambier, J. C., S. B. Gauld, K. T. Merrell, and B. J. Vilen. 2007. B-cell anergy: From transgenic models to naturally occurring anergic B cells? Nature Reviews Immunology 7:633–643. Carsetti, R., M. M. Rosado, and H. Wardmann, H. 2004. Peripheral development of B cells in mouse and man. Immunological Reviews 197:179–191. Casola, S. 2007. Control of peripheral B-cell development. Current Opinions in Immunology 19:143–149. Chung, J. B., R. A. Sater, M. L. Fields, J. Erikson, and J. G. Monroe. 2002. CD23 defines two distinct subsets of immature B cells which differ in their responses to T cell help signals. International Immunology 14:157–166. Dorshkind, K., and E. Montecino-Rodriguez. 2007. Fetal B-cell lymphopoiesis and the emergence of B-1-cell potential. Nature Reviews Immunology 7:213–219. Dzierzak, E., and N. A. Speck. 2008. Of lineage and legacy: The development of mammalian hematopoietic stem cells. Nature Immunology 9:129–136. Nemazee, D. 2006. Receptor editing in lymphocyte development and central tolerance. Nature Reviews Immunology 6:728–740. Nutt, S. L., and Kee, B. L. 2007. The transcriptional regulation of B cell lineage commitment. Immunity 26:715–725. Palis, J., et al. 2001. Spatial and temporal emergence of high proliferative potential hematopoietic precursors during murine embryogenesis. Proceedings of the National Academy of Sciences of the United States of America 98:4528–4533. Perez-Vera, P., A. Reyes-Leon, and E. M. Fuentes-Panana. 2011. Signaling proteins and transcription factors in normal and malignant early B cell development. Bone Marrow Research 2011:502,751. Pillai, S., and A. Cariappa. 2009. The follicular versus marginal zone B lymphocyte cell fate decision. Nature Reviews Immunology 9:767–777. Rajewsky, K., and H. von Boehmer. 2008. Lymphocyte development: Overview. Current Opinions in Immunology 20:127–130. Fuxa, M., and J. A. Skok. 2007. Transcriptional regulation in early B cell development. Current Opinions in Immunology 19:129–136. Ramirez, J., K. Lukin, and J. Hagman. 2010. From hematopoietic progenitors to B cells: Mechanisms of lineage restriction and commitment. Current Opinions in Immunology 22:177– 184. Hoek, K. L., et al. 2006. Transitional B cell fate is associated with developmental stage-specific regulation of diacylglycerol and calcium signaling upon B cell receptor engagement. Journal of Immunology 177:5405–5413. Srivastava, B., W. J. Quinn, 3rd, K. Hazard, J. Erikson, and D. Allman. 2005. Characterization of marginal zone B cell precursors. Journal of Experimental Medicine 202:1225– 1234. Kee, B. L. 2009. E and ID proteins branch out. Nature Reviews Immunology 9:175–184. Tokoyoda, K., T. Egawa, T. Sugiyama, B. I. Choi, and T. Nagasawa, T. 2004. Cellular niches controlling B lymphocyte behavior within bone marrow during development. Immunity 20:707–718. Koralov, S. B., et al. 2008. Dicer ablation affects antibody diversity and cell survival in the B lymphocyte lineage. Cell 132:860–874. Kurosaki, T., H. Shinohara, and Y. Baba. 2010. B cell signaling and fate decision. Annual Review of Immunology 28:21–55. Liao, D. 2009. Emerging roles of the EBF family of transcription factors in tumor suppression. Molecular Cancer Research 7:1893–1901. Mackay, F., and P. Schneider. 2009. Cracking the BAFF code. Nature Reviews Immunology 9:491–502. Malin, S., et al. 2010. Role of STAT5 in controlling cell survival and immunoglobulin gene recombination during pro-B cell development. Nature Immunology 11:171–179. Mikkola, H. K., and S. H. Orkin. 2006. The journey of developing hematopoietic stem cells. Development 133:3733–3744. Monroe, J. G., and K. Dorshkind. 2007. Fate decisions regulating bone marrow and peripheral B lymphocyte development. Advances in Immunology 95:1–50. Nagasawa, T. 2006. Microenvironmental niches in the bone marrow required for B-cell development. Nature Reviews Immunology 6:107–116. Tung, J. W., and L. A. Herzenberg. 2007. Unraveling B-1 progenitors. Current Opinions in Immunology 19:150–155. von Boehmer, H., and F. Melchers. 2010. Checkpoints in lymphocyte development and autoimmune disease. Nature Immunology 11:14–20. Whitlock, C. A., and O. Witte. 1982. Long-term culture of B lymphocytes and their precursors from murine bone marrow. Proceedings of the National Academy of Sciences of the United States of America 79:3608–3612. Xiao, C., and K. Rajewsky. 2009. MicroRNA control in the immune system: Basic principles. Cell 136:26–36. Yin, T., and L. Li. 2006. The stem cell niches in bone. Journal of Clinical Investigation 116:1195–1201. Useful Web sites www.bio.davidson.edu/courses/immunology/ Flash/Bcellmat.html An unusual animation of B-cell development. PA R T I V 356 S T U D Y | Adaptive Immunity: Development Q U E S T I O N S 1. You wish to study the development of B-1 B cells in the absence of the other two major B-cell subsets. You have a recipient Rag1/ mouse that you have already repopulated with T cells. What would you choose to be your source of B-1 progenitors and why? Which anatomical sites would you expect to harvest the B-1 B cells from? (b) 2. Describe the phenotypic and functional differences between T1 and T2 immature B cells. 3. Following expression of the pre-B-cell receptor on the pro- genitor B-cell surface, the B cell undergoes a few rounds of cell division. What purpose does this round of division serve in the development of the B-cell repertoire? 4. Immature B cells bearing potentially autoimmune recep- tors can be managed in three ways to minimize the probability of disease. Describe these three strategies, noting whether they are shared by T-cell progenitors. 5. You suspect that a new transcription factor is expressed at the pre-pro-B-cell stage of development. How would you test your hypothesis? What is the status of heavy-and lightchain rearrangement at this stage of development and how would you test it? 6. How would you determine whether a particular stage of B-cell development occurs in association with a stromal cell that expresses CXCL12? 7. Describe the order in which B-cell receptor genes undergo rearrangement, indicating at what steps you might expect to see the B cell express one or both chains on the cell surface. In what sense(s) does this gene rearrangement process mimic the analogous progression in  T cells, and in what ways do the two processes differ? ANALYZE THE DATA The two columns of data in the following figure below are flow cytometric plots that describe the levels of the antigens denoted on the x and y axes. The left column represents the antigens present on spleen (part A) and bone marrow (parts B) from wild-type (genetically normal) animals. The plots represent all lymphocytes in the spleen (part A) or B-cell progenitor and precursor cells in the bone marrow (part B). The right column shows the same plots from animals in which the Dicer gene has been knocked out. As you recall, the Dicer gene is required for the maturation of controlling miRNAs. a. For each pair of plots, describe the differences in the cell populations, indicating whether the differences reflect losses or gains in particular developing B-cell populations. b. At what point(s) in B-cell development do you think miRNAs are functioning? Wild type mouse (a) Dicer knockout mouse ANALYZE THE DATA The following figure is derived from the same paper as those above. In this case the data are expressed as histograms, in which the y axis represents the number of cells binding the molecule shown on the x axis, Annexin V. Annexin V binds to phosphatidyl serine on the outer leaflet of cell membranes. Phosphatidyl serine is found on the outer leaflet only in cells about to undergo apoptosis. The top two panels represent cells from a wild-type animal, and the bottom two panels represent cells from animals in which the Dicer gene has been knocked out. Pro-B cells Pre-B cells Wild type mouse Dicer knockout mouse a. Does the presence of Dicer have an effect on the frac- tion of pro-B cells undergoing apoptosis? Explain your reasoning. b. Does the presence of Dicer have an effect on the fraction of pre-B cells undergoing apoptosis? Again explain your reasoning. c. Describe one function that you now think miRNAs fulfill in B-cell development. 11 T-Cell Activation, Differentiation, and Memory T he interaction between a naïve T cell and an antigen-presenting cell (APC) is the initiating event of the adaptive immune response. Prior to this, the innate immune system has been alerted at the site of infection or tissue damage, and APCs, typically dendritic cells, have been activated via their pattern recognition receptors. These cells may have engulfed extracellular (or opsonized intracellular) pathogens, or they may have been infected by an intracellular pathogen. In either case, they have processed and presented peptides from these pathogens in complex with surface MHC class I and class II molecules, and have made their way to a local (draining) lymph node and/or the spleen. The APCs have taken up residence in the T-cell zones of the lymph node or spleen to join networks of other cells that are continually scanned by roving naïve CD8⫹ and CD4⫹ T cells, which recognize MHC class I-peptide and MHC class IIpeptide complexes, respectively. We have seen that each mature T cell expresses a unique antigen receptor that has been assembled via random gene rearrangement during T-cell development in the thymus (Chapter 9). Because developing T cells undergo selection events within the thymus, each mature, naïve T cell is tolerant to self antigens, and restricted to self-MHC (Chapter 9). Some naïve T cells have committed to the CD8⫹ cytotoxic T-cell lineage, some to the CD4⫹ helper T-cell lineage. When a naïve CD8⫹ or CD4⫹ T cell binds tightly to an MHC-peptide complex expressed by an activated dendritic cell, it becomes activated by signals generated through the TCR (see Chapter 3). These signals, in concert with signals from other factors that we will describe below, stimulate the T cell to proliferate and differentiate into an effector cell. As you know, naïve CD8⫹ T cells become cytotoxic cells in response to engagement of MHC class I-peptide combinations. Although we refer to the activation of CD8⫹ T cells in this chapter, we will discuss their effector functions in detail in Chapter 13. Naïve CD4⫹ T cells become helper cells in response to engagement of MHC class II-peptide combinations (Overview Figure 11-1). Dendritic cell (orange) interacting with T cells (green). [M. Rohde, HZI, Braunschweig, Germany.] ■ T-Cell Activation and the Two Signal Hypothesis ■ T-Cell Differentiation ■ T-Cell Memory This chapter focuses on this event, which is critical for the development of both humoral and cell-mediated immunity, as well as the development of B-cell and CD8⫹ T-cell memory. As discussed below, CD4⫹ T cells can differentiate into a surprising number of distinct helper subsets, each of which has a different function in combating infection. In this chapter, we briefly review the cellular and molecular events that activate T cells and then deepen your understanding of the costimulatory interactions that play an important role in determining the outcome of T cell-APC interactions. We then discuss the outcomes of naïve T-cell activation—the development of effector and memory T cells—focusing primarily on the different fates and functions of the CD4⫹ helper T-cell subsets that drive the adaptive response. Which helper subset a naïve CD4⫹ T cell becomes depends on the types of signals (e.g., cytokines, costimulatory signals) they receive from the dendritic cells they engage via their TCRs. And as described in Chapter 5, the signals dendritic cells are able to deliver depend in large part on the pathogen to which they have been exposed (see Figure 5-18). Investigators are still working to understand all the variables involved in determining the lineage choices of T helper cells, but we introduce you to the current thinking. 357 358 PA R T V | Adaptive Immunity: Effector Responses 11-1 OVERVIEW FIGURE T-Cell Activation and Differentiation Antigen recognition Clonal expansion Activation Differentiation Effector functions Effector CD4+ T cell Naïve CD4+ T cell Activation of macrophages, B cells, other cells Memory CD4+ T cell IL-2R IL-2 Cytokines APC Other cellular sources of cyotkines Naïve CD8+ T cell IL-2R IL-2 Lymphoid organs Activation of a naïve T cell in a secondary lymphoid organ results in the generation of effector and memory T cells. Activation requires several receptor-ligand interactions between the T cell and a dendritic cell, as well as signals through cytokines produced by the We also describe the known functions of the specialized helper cells, focusing on TH1, TH2, TH17, TFH, and TREG cells. Finally, we close the chapter with a discussion of T-cell memory, which is dependent on CD4⫹ T cell help, and describe both what is known and what is currently under investigation. A Classic Experiment box and Clinical Focus box are offered as a pair and describe the basic research behind the discovery of the costimulatory molecule CD28, an essential participant in naïve T-cell activation, and then the development of a molecular therapy for autoimmune diseases that takes advantage of what we know about the biology of costimulation. These boxes, together, illustrate the powerful connections between basic research and clinical development, which underlie translational research, an effort to bring bench scientific discovery to the “bedside” that has captured the imagination of many biomedical investigators. Effector CD8+ T cell (CTL) Memory CD8+ T cell Killing of infected “target cells”; macrophage activation Peripheral tissues activating APC, as well as other supportive cells in the lymphoid organ. Effector CD4⫹ T cells become helper T cells (TH) and secrete cytokines that enhance the activity of many other immune cells. Effector CD8⫹ T cells are cytotoxic cells (TC) that kill infected cells. The Advances box describes a more recent effort to figure out precisely how many T-cell receptors must be engaged to initiate T-cell activation. The answer was initially surprising, yet in hindsight may not be surprising at all. The final Clinical Focus box discusses how a disease, an “experiment of nature,” has helped us to better understand the basic biology and physiological function of the effector cells introduced in this chapter. T-Cell Activation and the Two-Signal Hypothesis CD4⫹ and CD8⫹ T cells leave the thymus and enter the circulation as resting cells in the G0 stage of the cell cycle. These naïve T cells are mature, but they have not yet encountered antigen. Their chromatin is condensed, they have very little cytoplasm, and they exhibit little transcriptional activity. T-Cell Activation, Differentiation, and Memory However, they are mobile cells and recirculate continually among the blood, lymph, and secondary lymphoid tissues, including lymph nodes, browsing for antigen. It is estimated that each naïve T cell recirculates from blood through lymph nodes and back again every 12 to 24 hours. Because only about 1 in 105 naïve T cells is likely to be specific for any given antigen, this large-scale recirculation increases the chances that a T cell will encounter appropriate antigen. If a naïve T cell does not bind any of the MHC-peptide complexes encountered as it browses the surfaces of stromal cells of a lymph node, it exits through the efferent lymphatics, ultimately draining into the thoracic duct and rejoining the blood (see Chapter 2). However, if a naïve T cell does encounter an APC expressing an MHC-peptide to which it can bind, it will initiate an activation program that produces a diverse array of cells that orchestrate efforts to clear infection. Recall from Chapter 3 that a successful T cell-APC interaction results in the stable organization of signaling molecules into an immune synapse (Figure 11-2). The TCR/ MHC-peptide complexes and coreceptors are aggregated in the central part of this synapse (central supramolecular activating complex, or cSMAC). The intrinsic affinity between the TCR and MHC-peptide surfaces is quite low (Kd ranges from 10⫺4 M to 10⫺7 M) and is stabilized by the activity of several molecules which together increase the avidity (the combined affinity of all cell-cell interactions) of the cellular interaction. The coreceptors CD4 and CD8, which are found in the cSMAC, stabilize the interaction between TCR and MHC by binding MHC class II and MHC class I molecules, respectively. Interactions between adhesion molecules and their ligands (e.g., LFA-1/ICAM-1 and CD2/LFA-3) help to sustain the signals generated by allowing long-term cell interactions. These molecules are organized around the central aggregate, forming the peripheral or “p” SMAC. However, even the increased functional avidity offered by coreceptors and adhesion molecules is still not sufficient to fully activate a T cell. Interactions between costimulatory receptors on T cells (e.g., CD28) and costimulatory ligands on dendritic cells (e.g., CD80/86) provide a second, required signal. In addition, as you will see below, a third set of signals, provided by local cytokines (Signal 3), directs T-cell differentiation into distinct effector cell types. Costimulatory Signals Are Required for Optimal T-Cell Activation and Proliferation What evidence pointed to a requirement for a second, costimulatory signal? In 1987, Helen Quill and Ron Schwartz recognized that, in the absence of functional APCs, isolated high affinity TCR-MHC interactions actually led to T-cell nonresponsiveness rather than activation—a phenomenon they called T-cell anergy. Their studies led to the simple but powerful notion that not one but two signals were required for full T-cell activation: Signal 1 is provided by antigen-specific TCR engagement (which can be enhanced by coreceptors and adhesion molecules), and Signal 2 is provided by contact with | CHAPTER 11 359 a costimulatory ligand, which can only be expressed by a functional APC. When a T cell receives both Signal 1 and Signal 2, it will be activated to produce cytokines that enhance entry into cell cycle and proliferation (Figure 11-3). It is now known that Signal 2 results from an interaction between specific costimulatory receptors on T cells and costimulatory ligands on dendritic cells (Table 11-1). Recall from Chapter 5 that dendritic cells and other APCs become activated by antigen binding to PRRs, to express costimulatory ligands (e.g., CD80 and CD86) and produce cytokines that enhance their ability to activate T cells. CD28 is the most commonly cited example of a costimulatory receptor, but other related molecules that provide costimulatory signals during T-cell activation have since been identified and are also described below. Because these molecules enhance TCR signaling, they are collectively referred to as “positive” costimulatory receptors and ligands. Negative costimulatory receptors, which inhibit TCR signaling, have also been identified. Although our understanding of their specific functions is incomplete, as a group these play important roles in (1) maintaining peripheral T-cell tolerance and (2) reducing inflammation both after the natural course of an infection and during responses to chronic infection. As you can imagine, the expression and activity of negative and positive costimulatory molecules must be carefully regulated temporally and spatially. Naïve T cells, for example, do not express negative costimulatory receptors, allowing them to be activated in secondary lymphoid tissue during the initiation of an immune response. On the other hand, effector T cells up-regulate negative costimulatory receptors at the end of an immune response, when proliferation is no longer advantageous. However, these generalizations belie the complexity of regulation of this highly important costimulatory network, and investigators are still working to understand the details. Below we introduce aspects of the structure, function, expression, and, when known, regulation of several positive and negative costimulators. Positive Costimulatory Receptors: CD28 CD28, a 44 kDa glycoprotein expressed as a homodimer, was the first costimulatory molecule to be discovered (see Classic Experiment Box 11-1). Expressed by all naïve and activated human and murine CD4⫹ T cells, all murine CD8⫹ T cells, and, interestingly, only 50% of human CD8⫹ T cells, it markedly enhances TCR-induced proliferation and survival by cooperating with T-cell receptor signals to induce expression of the pro-proliferative cytokine IL-2 and the prosurvival bcl-2 family member, bcl-xL. CD28 binds to two distinct ligands of the B7 family of proteins: CD80 (B7-1) and CD86 (B7-2). These are members of the immunoglobulin superfamily, which have similar extracellular domains. Interestingly, their intracellular regions differ, suggesting that they might not simply act as passive ligands; rather, they may have the ability to generate signals that influence the APC, a view that has some experimental support. PA R T V 360 | Adaptive Immunity: Effector Responses (a) cSMAC TCR/CD3 Coreceptors CD4 or CD8 Costimulatory receptors (CD28) pSMAC Adhesion molecules LFA-1/ICAM-1 LFA-3/CD2 (b) Antigen-presenting cell CD4+ T cell CD8+ T cell Antigen-presenting cell Antigen Antigen pSMAC Peptide LFA-3 CD2 LFA-3 Peptide TCR− CD3 Class II MHC p56lck CD4 Class I MHC LFA-1 CD80 or CD86 ICAM-1 FIGURE 11-2 Surface interactions responsible for T-cell activation. (a) A successful T-cell/dendritic-cell interaction results in the organization of signaling molecules into an immune synapse. A scanning electron micrograph (left) shows the binding of a T cell (artificially colored yellow) and dendritic cell (artificially colored blue). A fluorescent micrograph (right) shows a cross-section of the immune synapse, where the TCR is stained with fluorescein (green) and adhesion molecules (specifically LFA-1) are stained with phycoerythrin (red). Other molecules that can be found in the central part of the synapse (central supramolecular activation complex [cSMAC]) and the peripheral part of the synapse (pSMAC) are listed. (b) The interactions between a CD4⫹ (left) or CD8⫹ (right) T-cell and its activating CD28 LFA-1 SS CD8 SS CD28 pSMAC TCR− CD3 SS cSMAC SS CD2 CD80 or CD86 ICAM-1 dendritic cell. A dendritic cell (to the right of each diagram) can engulf an antigen and present peptide associated with MHC class II to a CD4⫹ T cell or can load internal peptides into MHC class I and present the combination to a CD8⫹ T cell. Binding of the TCR to MHCpeptide complexes is enhanced by the binding of coreceptors CD4 and CD8 to MHC class II and class I, respectively. CD28 interactions with CD80/86 provide the required costimulatory signals. Adhesion molecule interactions, two of which (LFA-1/ICAM-1, LFA-3/CD2) are depicted, markedly strengthen the connection between the T cell and APC or target cell so that signals can be sustained. [(a) right: Michael L. Dustin, J Clin Invest. 2002; 109(2): 155, Fig.1. doi:10.1172/JCI14842. Left: Dr. Olivier Schwartz, Institut Pasteur/Photo Researchers.] T-Cell Activation, Differentiation, and Memory Normal APC 1 TCR signaling Gene expression 2 Costimulatory interaction Autocrine (e.g., IL-2) 3 Cytokine signaling Paracrine (e.g., IL-12) FIGURE 11-3 Three signals are required for activation of a naïve T cell. The TCR/MHC-peptide interaction, along with CD4 and CD8 coreceptors and adhesion molecules, provide Signal 1. Costimulation by a separate set of molecules, including CD28 (or ICOS, not shown) provide Signal 2. Together, Signal 1 and Signal 2 initiate a signal transduction cascade that results in activation of transcription factors and cytokines (Signal 3) that direct T-cell proliferation (IL-2) and differentiation (polarizing cytokines). Cytokines can act in an autocrine manner, by stimulating the same cells that produce them, or in a paracrine manner, by stimulating neighboring cells. Although most T cells express CD28, most cells in the body do not express its ligands. In fact, only professional APCs have the capacity to express CD80/86. Mature dendritic cells, the best activator of naïve T cells, appear to con- TABLE 11-1 | CHAPTER 11 361 stitutively express CD80/86, and macrophages and B cells have the capacity to up-regulate CD80/86 after they are activated by an encounter with pathogen (see Chapter 5). Positive Costimulatory Receptors: ICOS Since the discovery of CD28, several other structurally related receptors have been identified. Like CD28, the closely related inducible costimulator (ICOS) provides positive costimulation for T-cell activation. However, rather than binding CD80 and CD86, ICOS binds to another member of the growing B7 family, ICOS-ligand (ICOS-L), which is also expressed on a subset of activated APCs. Differences in patterns of expression of CD28 and ICOS indicate that these positive costimulatory molecules play distinct roles in T-cell activation. Unlike CD28, ICOS is not expressed on naïve T cells; rather, it is expressed on memory and effector T cells. Investigations suggest that CD28 plays a key costimulatory role during the initiation of activation and ICOS plays a key role in maintaining the activity of already differentiated effector and memory T cells. Negative Costimulatory Receptors: CTLA-4 The discovery of CTLA-4 (CD152), the second member of the CD28 family to be identified, caused a stir. Although closely related in structure to CD28 and also capable of binding both CD80 and CD86, CTLA-4 did not act as a positive costimulator. Instead, it antagonized T-cell activating signals and is now referred to as a negative costimulatory receptor. CTLA-4 is not expressed constitutively on resting T cells. Rather, it is induced within 24 hours after activation of a T-cell costimulatory molecules and their ligands Costimulatory receptor on T cell Costimulatory ligand Activity CD28 CD80 (B7-1) or CD86 (B7-2) Expressed by professional APCs, (and medullary thymic epithelium) Activation of naïve T cells ICOS ICOS-L Expressed by B cells, some APCs, and T cells Maintenance of activity of differentiated T cells; a feature of T-/B-cell interactions CTLA-4 CD80 (B7-1) or CD86 (B7-2) Expressed by professional APCs (and medullary thymic epithelium) Negative regulation of the immune response (e.g., maintaining peripheral T-cell tolerance; reducing inflammation; contracting T-cell pool after infection is cleared) PD-1 PD-L1 or PD-L2 Expressed by professional APCs, some T and B cells, and tumor cells Negative regulation of the immune response, regulation of TREG differentiation BTLA HVEM Expressed by some APCs, T and B cells Negative regulation of the immune response, regulation of TREG differentiation (?) Positive costimulation Negative costimulation 362 PA R T V | Adaptive Immunity: Effector Responses CLASSIC EXPERIMENT Discovery of the First Costimulatory Receptor: CD28 In 1989, Navy immunologist Carl June took the first step toward filling (and revealing) the considerable gaps in our understanding of T-cell proliferation and activation by introducing a new actor: CD28. CD28 had been recently identified as a dimeric glycoprotein expressed on all human CD4⫹ T cells and half of human CD8⫹ T cells, and preliminary data suggested that it enhanced T-cell activation. June and his colleagues specifically wondered if CD28 might be related to the Signal 2 that was known to be provided by APCs (sometimes referred to as “accessory cells” in older literature). June and his colleagues isolated T cells from human blood by density gradient centrifugation and by depleting a T-cellenriched population of cells that did not express CD28 (negative selection). They then measured the response of these CD28⫹ T cells to TCR stimulation in the presence or absence of CD28 engagement (Figure 1a). To mimic the TCR-MHC interaction, June and colleagues used either monoclonal antibodies to the CD3 complex or the mitogen phorbol myristyl acetate (PMA), a protein kinase C (PKC) activator. To engage the CD28 molecule, they used an anti-CD28 monoclonal antibody. They included two negative controls: one population that was cultured in growth medium with no additives and another population that was exposed to phytohaemagglutinin (PHA), which was known to activate T cells only in the presence of APCs. This last control was clever and would be used to demonstrate that the researchers’ isolated populations were not contaminated with APCs, which would express their own sets of ligands and confound interpretation. June and colleagues measured the proliferation of each of these populations by measuring incorporation of a radioactive (tritiated) nucleotide, [3H]Uridine, which is incorporated by cells that are synthesizing new RNA (and, hence, are showing signs of activation). Their results were striking, particularly when responses to PMA were examined (Figure 1b). As expected, T cells grown without stimulation or with incomplete stimulation (PHA) naïve T cell and peaks in expression within 2 to 3 days poststimulation. Peak surface levels of CTLA-4 are lower than peak CD28 levels, but because it binds CD80 and CD86 with markedly higher affinity, CTLA-4 competes very favorably with CD28. Interestingly, CTLA-4 expression levels increase in proportion to the amount of CD28 costimulation, suggesting that CTLA-4 acts to “put the brakes on” the proproliferative influence of TCR-CD28 engagement. The importance of this inhibitory function is underscored by the phenotype of CTLA-4 knockout mice, whose T cells proliferate without control, leading to lymphadenopathy (greatly enlarged lymph nodes), splenomegaly (enlarged spleen), autoimmunity, and death within 3 to 4 weeks after birth. Negative Costimulatory Receptors: PD-1 and BTLA PD-1 (CD279) and B and T lymphocyte attenuator (BTLA (CD272)) are relatively new additions to our list of negative costimulatory receptors. Although more distantly related to CD28 family members than CTLA-4, they also inhibit TCRmediated T-cell activation. program death-1 (PD-1) is expressed by both B and T cells and binds to two ligands, remained quiescent, exhibiting no nucleotide uptake. Cell groups treated with stimuli that were known to cause T-cell proliferation—anti-CD3 and PMA— showed evidence of activation, with CD3 engagement producing relatively more of a response than PMA at the time points examined. The cells treated with anti-CD28 only were just as quiescent as the negative control samples, indicating that engagement of CD28, alone, could not induce activation. However, when CD28 was engaged at the same time cells were exposed to PMA, incorporation of [3H]Uridine increased markedly. T cells cotreated with anti-CD28 and anti-CD3 also took up more [3H]Uridine than those treated with anti-CD3 alone. What new RNA were these cells producing? Using Northern blot analysis and functional assays (bioassays) to characterize the cytokines in culture supernatant of the stimulated cells, June and colleagues went on to show that CD28 stimulation induced anti-CD3 stimulated T cells to produce higher levels of cytokines PD-L1 (B7-H1) and PD-L2 (B7-DC), which are also members of the CD80/86 family. PD-L2 is expressed predominantly on APCs; however, PD-L1 is expressed more broadly and may help to mediate T-cell tolerance in nonlymphoid tissues. Recent data suggest that PD-L1/PD-L1/2 interactions regulate the differentiation of regulatory T cells. BTLA is more broadly expressed: not only has it been found on conventional TH cells, as well as ␥␦ T cells and regulatory T cells, but it is also expressed on NK cells, some macrophages and dendritic cells, and most highly on B cells. Interestingly, BTLA’s primary ligand appears not to be a B7 family member, but a TNF receptor family member known as herpesvirus-entry mediator (HVEM), which is also expressed on many cell types. Studies on the role of this interesting costimulatory receptor-ligand pair are ongoing, but there are indications that BTLA-HVEM interactions also play a role in down-regulating inflammatory and autoimmune responses. As the genome continues to be explored, additional costimulatory molecules—both negative and positive in influence—are likely to be identified. Understanding their T-Cell Activation, Differentiation, and Memory | CHAPTER 11 363 BOX 11-1 10 PMA or anti-CD3 1 2 Anti-CD28 FIGURE 1 (a) June’s experimental setup used anti-CD3 monoclonal antibodies or a mitogen, PMA, to provide Signal 1, and anti-CD28 antibodies to provide Signal 2. (b) June measured [3H]Uridine incorporation, an indicator of RNA synthesis, in response to various treatments. Addition of stimulating antiCD28 antibody increased RNA synthesis in response to activation by PMA or anti-CD3. [Part (b) adapted from C. Thompson et al., 1989, CD28 activation pathway 3H-Uridine Gene expression incorporation (CPMx10-3) 9 8 7 6 5 4 3 2 1 0 MED PHA regulates the production of multiple T-cell-derived lymphokines/cytokines. Proceedings of the National Academy of Sciences of the United States of America 86:1333.] involved in antiviral, anti-tumor, and proliferative activity, generating an increase in T-cell immune response. When this paper was published, June did not know the identity of the natural ligand for CD28, or even if the homodimer could be activated in a natural immune context. We now know that CD28 binds to CD80/86 (B7), providing the critical Signal 2 during naïve T-cell activation, a signal required for optimal up-regulation of IL-2 and the IL-2 receptor. Finding this second switch capable of modulating T-cell activation was only the beginning of a landslide of discoveries of additional costimulatory signals—positive and negative—involved in T-cell activation, and the recognition that T cells are regulation and function will continue to occupy the attention of the immunological community and has already provided the clinical community with new tools for manipulating the immune response during transplantation and disease (see Clinical Focus Box 11-2). Clonal Anergy Results if a Costimulatory Signal Is Absent Experiments with cultured cells show that if a resting T cell’s TCR is engaged (Signal 1) in the absence of a suitable costimulatory signal (Signal 2), that T cell will become unresponsive to subsequent stimulation, a state referred to as anergy. There is good evidence that both CD4⫹ and CD8⫹ T cells can be anergized, but most studies of anergy have been conducted with CD4⫹ TH cells. Anergy can be demonstrated in vitro with systems designed to engage the TCR in the absence of costimulatory molecule engagement. For instance, T cells specific for a MHC-peptide complex can be induced to proliferate in vitro by incubation with activated APCs that express both the αCD28 PMA PMA + αCD28 αCD3 αCD3 + αCD28 even more subtly perceptive than we once appreciated. Based on a contribution by Harper Hubbeling, Haverford College, 2011. Thompson, C., et al. 1989. CD28 activation pathway regulates the production of multiple T-cell-derived lymphokines/cytokines. Proceedings of the National Academy of Sciences of the United States of America 86(4):1333–1337. appropriate MHC-peptide combination and CD80/86. However, glutaraldehyde-fixed APCs, which express MHC class II-peptide complexes, but cannot be induced to express CD80/86, render T cells unresponsive (Figure 11-4a). These anergic T cells are no longer able to secrete cytokines or proliferate in response to subsequent stimulation (Figure 11-4b). T-cell anergy can also be induced in vitro by incubating T cells with normal, CD80/86-expressing APCs in the presence of the Fab portion of anti-CD28, which blocks the interaction of CD28 with CD80/86 (Figure 11-4c). In vivo, the requirement for costimulatory ligands to fully activate a T cell decreases the probability that circulating autoreactive T cells will be activated and become dangerous. For instance, a naïve T cell expressing a T-cell receptor specific for an MHC class I insulin-peptide complex would be rendered nonresponsive if it encountered a ␤-islet cell expressing this MHC class I-peptide complex. Why? ␤-islet cells cannot be induced to express costimulatory ligands, and the encounter would result in T-cell anergy, preventing an immune attack on these insulinproducing cells. 364 PA R T V | Adaptive Immunity: Effector Responses BOX 11-2 CLINICAL FOCUS Costimulatory Blockade Taking an idea from the research setting to a therapeutic reality—from “bench to bedside”—is a dream for many, but a reality for few. Immunologists recognized the therapeutic promise of costimulatory receptors as soon as they were discovered. Reagents that would block these interactions could block the activation of destructive T cells, which were known to be responsible for many autoimmune disorders and for transplantation rejection. Several investigators specifically recognized that soluble versions of CD28 and CTLA-4 could be very valuable. By blocking the interaction between costimulatory receptors and their CD80/86 ligands, soluble CD28 and CTLA-4 could inhibit destructive T-cell responses (e.g., those involved in autoimmune disease or transplantation rejection). Couple this idea with technological advances in protein design and you had a novel reagent—and a potential drug. In the early 1990s, at least two groups converted human CTLA-4 into a soluble protein by fusing the extracellular domain of human CTLA-4 with the Fc portion of an IgG1 antibody (Figure 1). The Fc portion enhances (1) the ability to manipulate a protein during production by taking advantage of Fc binding as well as (2) the distribution of the reagent in an organism (the Fc portion confers some of the antibodies’ tissue distribution behaviors). The Fc portion is modified so that it does not CTLA4 IgG1 domain FIGURE 1 Structure of CTLA-4 Ig The extracellular domain of human CTLA-4 is fused to a modified Fc region of human IgG1. [PDB IDs 1IGY and 3OSK.] bind to Fc receptors and cause unintended cytotoxicity. Using this new protein, Peter Linsley et al. found that their soluble CTLA4-Ig bound CD80/86 with higher affinity than CD28-Ig, and could therefore more potently block costimulation. The development of CTLA-4 Ig as a drug did not proceed without difficulty. Originally designed and tested as a treatment for T-cell-mediated transplantation rejection, it did not originally perform as well as expected. However, hope in its utility was revived when it showed significant promise as a treatment for rheumatoid arthritis. The marketing of CTLA-4 Ig was not without controversy, either. At least two groups claimed patent rights, and communication between companies Interactions between negative costimulatory receptors and ligands can also induce anergy. This phenomenon, which would typically only apply to activated T cells that have up-regulated negative costimulatory receptors, could help curb T-cell proliferation when antigen is cleared. Unfortunately, negative costimulation may also contribute to the T cell “exhaustion” during chronic infection, such as that caused by mycobacteria, HIV, hepatitis virus, and more. T cells specific for these pathogens express high levels of PD-1 and BTLA, and are functionally anergic. Some recent therapies, in fact, are designed to block this interaction and allow T cells to reactivate. and basic researchers was not always smooth. In late 2005, the FDA approved the use of CTLA-4 Ig (abatacept) for rheumatoid arthritis (RA). As of 2012 it is marketed as Orencia by Bristol Meyers Squibb, which shares patent royalties with Repligen Corporation. Abatacept shows promise in delaying the joint damage seen with RA, and clinical trials are also underway to evaluate its potential to ameliorate psoriasis, lupus, Type 1 diabetes, and more. This true bench-to-bedside story is still not finished and continues to be informed by basic researchers’ growing knowledge of the complexities of costimulation. Investigators have already developed a modified version of CTLA-4 Ig that differs in two amino acids in the CD80/86 binding region and exhibits a higher affinity for CD86, which may play a more important role in transplantation rejection. This drug is also showing promise in clinical trials and may result in fewer side effects than standard immunosuppressant therapies (e.g., cyclosporine), which are not specific for T cells. It is also important to recognize that CTLA-4 Ig not only blocks positive costimulatory reactions, but also has the potential to inhibit negative ones and in some circumstances could lead to enhancement of T-cell activity. Researchers will continue to draw from basic and clinical knowledge to determine how best to use the drug and to improve its design for enhanced safety and efficacy. Although anergy is a well-established phenomenon, the precise biochemical pathways that regulate this state of nonresponsiveness are still not fully understood. During the past few years, microarray analyses (see Chapter 20) have identified several key enzymes expressed by anergic T cells, including ubiquitin ligases that appear to target key components of the TCR signaling pathway for degradation by the proteasome. Cytokines Provide Signal 3 We have now seen that naïve T cells will initiate activation when they are costimulated by engagement with both T-Cell Activation, Differentiation, and Memory (a) 1 Fixed APC (no CD80/86) Anergic genes | CHAPTER 11 365 As we will see below, Signal 3 is also supplied by other cytokines (produced by APCs, T cells, NK cells, and others), known as polarizing cytokines, that play central roles not just in enhancing proliferation, but also in determining what types of effector cells naïve T cells will become. Antigen-Presenting Cells Have Characteristic Costimulatory Properties (b) Anergic T cell Normal APC No response (c) 1 Anergic genes Normal APC Fab anti-CD28 CD80/86 FIGURE 11-4 Experimental demonstration of clonal anergy versus clonal expansion. (a) Only Signal 1 is generated when resting T cells are incubated with glutaraldehyde-fixed APCs, which cannot be stimulated to up-regulate the costimulatory ligand CD80/86. (b) Anergic cells cannot respond to subsequent challenge, even when APCs can engage costimulatory receptors. (c) Anergy can also be induced when naïve T cells are incubated with normal APCs in the presence of the Fab portion of anti-CD28, which blocks interaction with CD80/86. MHC-peptide complexes and costimulatory ligands on dendritic cells. However, the consequences and extent of T-cell activation are also critically shaped by the activity of soluble cytokines produced by both APCs and T cells. These assisting cytokines are referred to, by some, as Signal 3 (see Figure 11-3). Cytokines bind surface cytokine receptors, stimulating a cascade of intracellular signals that can enhance both proliferation and/or survival (see Chapter 4). IL-2 is one of the best-known cytokines involved in T-cell activation and plays a key role in inducing optimal T-cell proliferation, particularly when antigen and/or costimulatory ligands are limiting. Costimulatory signals induce transcription of genes for both IL-2 and the ␣ chain (CD25) of the high-affinity IL-2 receptor. Signals also enhance the stability of the IL-2 mRNA. The combined increase in IL-2 transcription and improved IL-2 mRNA stability results in a 100-fold increase in IL-2 production by the activated T cell. Secretion of IL-2 and its subsequent binding to the high-affinity IL-2 receptor induces activated naïve T cells to proliferate vigorously. Which cells are capable of providing both Signal 1 and Signal 2 to a naïve T cell? Although almost all cells in the body express MHC class I, only professional APCs—dendritic cells, activated macrophages, and activated B cells—express the high levels of class II MHC molecules that are required for T-cell activation. Importantly, these same professional APCs are among the only two cell types capable of expressing costimulatory ligands. (The only other cell type known to have this capacity is the thymic epithelial cell; see Chapter 9.) Professional APCs are more diverse in function and origin than originally imagined, and each subpopulation differs both in the ability to display antigen and in the expression of costimulatory ligands (Figure 11-5). In the early stages of an immune response in secondary lymphoid organs, T cells encounter two main types of professional APCs: the dendritic cell and the activated B cell. Mature dendritic cells that have been activated by microbial components via their pattern recognition receptors (PRRs) are present throughout the T-cell zones. They express antigenic peptides in complex with high levels of MHC class I and II molecules. They also express costimulatory ligands and are the most potent activators of naïve CD4⫹ and CD8⫹ T cells. Resting B cells residing in the follicles gain the capacity to activate T cells after they bind antigen through their B-cell receptor (BCR). This engagement stimulates the up-regulation of MHC class II and costimulatory CD80/86, allowing the B cell to present antigen to CD4⫹ T cells they encounter at the border between the follicle and T-cell zone. Because of their unique ability to internalize antigen (e.g., pathogens) via specific BCRs and present them in MHC class II, B cells are most efficient at activating CD4⫹ T cells that are specific for the same pathogen for which they are specific. This situation serves the immune response very well, focusing the attention of antigen-specific CD4⫹ T cells activated in the T-cell zone on B cells activated by the same antigen in the neighboring follicle. The pairing of B cells with their helper T cells occurs at the junction between the B- and T-cell zones (see Chapter 14) and allows T cells to deliver the help required for B-cell proliferation, differentiation, and memory generation (see Chapter 12). Macrophages are also found in secondary lymphoid organs, but can activate T cells in a wide range of other peripheral tissues. They also must be activated to reveal their full antigenpresenting capacity. They up-regulate MHC molecules and costimulatory ligands in response to interactions with pathogens, as well as in response to cytokines produced by other cells, including IFN-␥. PA R T V 366 | Adaptive Immunity: Effector Responses Dendritic cell Resting B lymphocyte Macrophage Activated Activated Resting Activated Resting Class I MHC PAMPs and cytokines PAMPs Class I MHC Class I MHC Ag T cell help (IFN-γ) CD80/86 Antigen-presenting cell Class II MHC Class I Class I Class II MHC MHC CD80/86 MHC Class II MHC Class II MHC CD80/86 Dendritic cell Macrophage B cell Antigen uptake Endocytosis Phagocytosis Phagocytosis Receptor-mediated endocytosis Activation Mediated by pattern recognition receptors Mediated by pattern recognition receptors and enhanced by T-cell help Mediated by antigen recognition MHC Class II expression Increases with activation (may express low levels constitutively) Increases with activation Increases with activation (expresses low levels of constitutively) Costimulatory activity Up-regulation of CD80/86 with activation Up-regulation of CD80/86 with activation Up-regulation of CD80/86 with activation T-cell activation Naïve T cells Effector T cells Memory T cells Effector T cells Memory T cells Effector T cells Memory T cells Location Resting: Circulation peripheral tissues Activated: SLOs (T-cell zones) Tertiary tissues Resting: Circulation peripheral tissues Activated: SLOs (subcapsular cortex of lymph node, marginal zones of spleen) Peripheral tissues Resting: Circulation SLOs (follicles) Activated: SLOs (B cell/T-cell zone interface, germinal centers, and marginal zones) FIGURE 11-5 Differences in the properties of professional antigen-presenting cells that induce T-cell activation. This figure describes general features of three major classes of professional APCs. Dendritic cells are the best activators of naïve T cells. This may be due, in part, to relatively high levels of expression of MHC and costimulatory molecules when they are mature and activated. Activated B cells interact most efficiently with differentiated TH cells that are specific for the same antigen that activated them. Macrophages play several different roles, processing and distributing antigen in second- It turns out that there are several different dendritic cell and macrophage subtypes; however, their functions are still being clarified. Some are likely to activate or induce differentiation of specific effector T cells that travel to the site of infection, and some may be involved in reactivating memory cells that reside in tissues. Others may help to quell, rather than to activate, responses. Full understanding awaits more research. ary lymphoid tissues (SLOs) as well as interacting with effector cells in the periphery. It is important to recognize that the distinctions shown are rules of thumb only. Functions among the APC classes overlap, and the field now recognizes different subsets within each major group of APC, each of which may act independently on different T-cell subsets. This diversity may be a consequence of activation by different innate immune receptors or may reflect the existence of independent cell lineages. Note that activation of effector and memory T cells is not as dependent on costimulatory interactions. Superantigens Are a Special Class of T-Cell Activators Superantigens are viral or bacterial proteins that bind simultaneously to specific V␤ regions of T-cell receptors and to the ␣ chain of class II MHC molecules. V␤ regions are encoded by over 20 different V␤ genes in mice and 65 different genes in humans. Each superantigen displays a T-Cell Activation, Differentiation, and Memory TH cell Vβ β α Peptide for which TCR is not specific Superantigen Endogenous superantigen is membrane bound TCR α β Class II MHC APC FIGURE 11-6 Superantigen-mediated cross-linkage of Tcell receptor and class II MHC molecules. A superantigen binds to all TCRs bearing a particular V␤ sequence regardless of their antigen specificity. Exogenous superantigens are soluble secreted bacterial proteins, including various exotoxins. Endogenous superantigens are membrane-embedded proteins produced by certain viruses; they include Mls antigens encoded by mouse mammary tumor virus. “specificity” for one of these V␤ versions, which can be expressed by up to 5% of T cells, regardless of their antigen specificity. This clamp-like connection mimics a strong TCR-MHC interaction and induces activation, bypassing the need for TCR antigen specificity (Figure 11-6). Superantigen binding, however, does not bypass the need for costimulation; professional APCs are still required for full T-cell activation by these microbial proteins. TABLE 11-2 | CHAPTER 11 367 Both endogenous superantigens and exogenous superantigens have been identified. Exogenous superantigens are soluble proteins secreted by bacteria. Among them are a variety of exotoxins secreted by Gram-positive bacteria, such as staphylococcal enterotoxins, toxic shock syndrome toxin, and exfoliative dermatitis toxin. Each of these exogenous superantigens binds particular V␤ sequences in T-cell receptors (Table 11-2) and cross-links the TCR to a class II MHC molecule. Endogenous superantigens are cell-membrane proteins generated by specific viral genes that have integrated into mammalian genomes. One group, encoded by mouse mammary tumor virus (MTV), a retrovirus that is integrated into the DNA of certain inbred mouse strains, produces proteins called minor lymphocyte-stimulating (Mls) determinants, which bind particular V␤ sequences in T-cell receptors and cross-link the TCR to class II MHC molecules. Four Mls superantigens, originating from distinct MTV strains, have been identified. Because superantigens bind outside the TCR antigenbinding cleft, any T cell expressing that particular V␤ sequence will be activated by a corresponding superantigen. Hence, the activation is polyclonal and can result in massive T-cell activation, resulting in overproduction of TH-cell cytokines and systemic toxicity. Food poisoning induced by staphylococcal enterotoxins and toxic shock induced by toxic shock syndrome toxin are two examples of disorders caused by superantigen-induced cytokine overproduction. Given that superantigens result in T-cell activation and proliferation, what adaptive value could they possibly have for the pathogens that make them? The answer is not clear, but there is evidence that such antigen-nonspecific T-cell Exogenous superantigens and their V specificity V SPECIFICITY Superantigen * Disease Mouse Human SEA Food poisoning 1, 3, 10, 11, 12, 17 nd SEB Food poisoning 3, 8.1, 8.2, 8.3 3, 12, 14, 15, 17, 20 Staphylococcal enterotoxins SEC1 Food poisoning 7, 8.2, 8.3, 11 12 SEC2 Food poisoning 8.2, 10 12, 13, 14, 15, 17, 20 SEC3 Food poisoning 7, 8.2 5, 12 SED Food poisoning 3, 7, 8.3, 11, 17 5, 12 Food poisoning 11, 15, 17 5.1, 6.1–6.3, 8, 18 Toxic shock syndrome toxin (TSST1) SEE Toxic shock syndrome 15, 16 2 Exfoliative dermatitis toxin (ExFT) Scalded skin syndrome 10, 11, 15 2 Mycoplasma arthritidis supernatant (MAS) Arthritis, shock 6, 8.1–8.3 nd Streptococcal pyrogenic exotoxins (SPE-A, B, C, D) Rheumatic fever, shock nd nd * Disease results from infection by bacteria that produce the indicated superantigens. 368 PA R T V | Adaptive Immunity: Effector Responses ADVANCES How Many TCR Complexes Must Be Engaged to Trigger T-Cell Activation? Until single-cell imaging techniques were developed, indirect methods were used to calculate how many ligands a T cell must recognize in order to be activated. Mark Davis and colleagues at Stanford School of Medicine approached this question using an acutely sensitive, twopart microscopic visualization technique (Figure 1). APCs were cultured (pulsed) briefly with peptide bound to a biotin molecule. When APCs are exposed to soluble peptides a small number will exchange it with a peptide bound to an MHC molecule on the cell surface. The number of peptides that actually bound to MHC could be determined in this system by adding a fluorescent (phycoerythrin) streptavidin conjugate, which binds the biotin of the biotinylated peptide and can be detected and quantified by fluorescent microscopy. The response of T cells specific for the complex could also be quantified using another fluorescent tool: fura-2, a dye that can enter cells and fluoresces when intracellular calcium is released. By adding fura-2 “loaded” T cells to the APCs bound to varying numbers of red fluorescent peptide (see Figure 1a), researchers could determine (1) if a T cell was activated and (2) how many molecules of peptide were present at the point of contact between the T cell and the APC—in other words, how many TCR/ MHC-biotinylated peptide complexes were required to stimulate the release of intracellular calcium. One set of data from these experiments is illustrated in Figure 1b. An activated, fura-2-loaded T cell (blue) is shown interacting with an APC in the upper left. The fluorescent micrograph of the peptide at the junction of the T cell and APC is shown in the top right image. The intensity of the red fluorescence varies with the number of peptides bound. (The image is “stretched” artificially because of the computer program used to quantify fluores- activation and inflammation hampers the development of a coordinated antigen-specific response that would most effectively thwart infection. Some speculate that the largescale proliferation and cytokine production that occur harm the cells and microenvironments that are required to start a normal response; others argue that these events induce T-cell tolerance to the pathogen. T-Cell Differentiation How does an interaction between a naïve T cell and a dendritic cell result in the generation of cells with effector functions? An activating, costimulatory interaction between a naïve T cell and an APC typically lasts 6 to 8 hours, a period that permits the development of a cascade of signaling events (see Chapter 3) that alter gene programs and induce differentiation into a variety of distinct effector and memory cell subtypes. Just a few TCR-MHC interactions (as described in Advances Box 11-3) stimulates a signaling cascade that, cence.) The fluorescence intensity calculated from this particular image indicated that only a single MHC-peptide combination was at the T cell-APC synapse. The plot below these images shows the intensity of fura-2 fluorescence (i.e., the increase in intracellular Ca2⫹) over time after T cell-APC engagement. The initial spike is an indicator that this single MHC-peptide could inspire robust Ca2⫹ signals. The investigators quantified many interactions in this way and definitively concluded that a single MHC-peptide combination could stimulate significant Ca2⫹ release. Maximal Ca2⫹ release was achieved in CD4⫹ T cells when as few as 10 TCR complexes were engaged. Similar results were obtained with CD8⫹ T cells. Irvine, D. J., M. A. Purbhoo, M. Krogsgaard, and M. M. Davis. 2002. Direct observation of ligand recognition by T cells. Nature 419:845–849. in combination with costimulation and signals received by soluble cytokines, culminate in the activation of “effector” molecules that regulate (1) cell survival, (2) cell cycle entry, and, as we shall see below, (3) cell differentiation. In 1 to 2 days after successful engagement with a dendritic cell in the T-cell zone of a secondary lymphoid organ, a naïve T cell will enlarge into a blast cell and begin undergoing repeated rounds of cell division. Signals 1 plus 2 induce up-regulation of expression and activity of prosurvival genes (e.g., bcl-2), as well as the transcription of genes for both IL-2 and the ␣ chain (CD25) of the highaffinity IL-2 receptor (Figure 11-7). The combined effect on a naïve T cell is activation and robust proliferation. Activated T cells divide 2 to 3 times per day for 4 to 5 days, generating a clone of progeny cells that differentiate into memory and effector T-cell populations. Activated T cells and their progeny gain unique functional abilities, becoming effector helper or cytotoxic T cells that indirectly and directly act to clear pathogen. CD8⫹ cytotoxic T cells leave the secondary lymphoid tissues and circulate T-Cell Activation, Differentiation, and Memory | CHAPTER 11 369 BOX 11-3 (a) T cell (b) T Ca2+ APC 4 PE snapshot 3 Ca2+ signal APC 2 1 0 -5 0 5 10 15 20 Time (minutes) FIGURE 1 Engagement of a single T-cell receptor can induce Ca2ⴙ signals in a T cell. (a) The experimental approach taken by the investigators to determine the ability of a single TCR-MHC interaction to generate Ca2⫹ signals. See text for details. (b) One example of original data on Ca2⫹ signaling generated from a single T cell that has engaged a single MHCpeptide ligand on the surface of an APC. The interaction between the T cell and APC is captured by bright field microscopy (top left) and by highresolution fluorescence microscopy (top right), where the arrow points to the single TCR/MHC-peptide interaction. The Ca2⫹ signal generated within the T cell over a 20-minute period is depicted in the graph and measured using software that quantifies the intensity of fura-2 fluorescence, which increases with a rise in cytosolic Ca2⫹ levels. [Irvine, D.J. et al. 2002 Nature 419:845–849.] to sites of infection where they bind and kill infected cells. CD4⫹ helper T cells secrete cytokines that orchestrate the activity of several other cell types, including B cells, macrophages, and other T cells. Some CD4+ T cells, particularly those that “help” B cells and those that generate lymphocyte memory, stay within secondary lymphoid tissue to continue to regulate the generation of the response. Others return to the sites of infection and enhance the activity of macrophages and cytotoxic cells. Effector cells tend to be short-lived and have life spans that range from a few days to a few weeks. However, the progeny of an activated T cell can also become long-lived memory T cells that reside in secondary and tertiary tissues for significant periods of time, providing protection against a secondary infection. Effector T cells come in more varieties than was originally anticipated, and each subset plays a specific and important role in the immune response. The first effector cell distinction to be recognized was between CD8⫹ T cells and CD4⫹ T cells. Activated CD8⫹ T cells acquire the ability to induce the death of target cells, becoming “killer” or “cytotoxic” T lymphocytes (CTL, or TC). Because cytotoxic CD8⫹ T cells recognize peptide bound to MHC class I, which is expressed by almost all cells in the body, they are perfectly poised to clear the body of cells that have been internally infected by the pathogen that resulted in their activation. On the other hand, activated CD4⫹ T cells (helper T cells or TH) acquire the ability to secrete factors that enhance the activation and proliferation of other cells. Specifically, they regulate activation and antibody production of B cells, enhance the phagocytic, anti-microbial, cytolytic, and antigen-presenting capacity of macrophages, and are required for the development of B-cell and CD8⫹ T-cell memory. As immunologists developed and adopted more tools to distinguish proteins expressed and secreted by helper T cells, it became clear that CD4⫹ helper T cells were particularly diverse, differentiating into several different subtypes, each of which secretes a signature set of cytokines. The cytokines secreted by CD4⫹ TH cells can either act directly on the same cell that produced them (acting in an autocrine fashion) or 370 PA R T V | Adaptive Immunity: Effector Responses Normal APC IL-2 IL-2 gene IL-2R gene 1 2 CD28 CD80/86 Activation IL-2 IL-2 receptor G1 M S G2 Several divisions M M E E E E Population of memory and effector cells FIGURE 11-7 Activation of a naïve T cell up-regulates expression of IL-2 and the high-affinity IL-2 receptor. Signal 1 and Signal 2 cooperate to enhance transcription and stability of mRNA for IL-2 and IL-2R. Secreted IL-2 will bind the IL-2R, which generates signals that enhance the entry of the T cell into the cell cycle and initiates several rounds of proliferation. Most of the cells differentiate into effector helper or cytotoxic cells, but some will differentiate into effector or central memory cells. can bind to receptors and act on cells in the vicinity (acting in a paracrine fashion). Below we describe the major features and functions of the best-characterized CD4⫹ helper T-cell subsets. Although CD8⫹ cytotoxic T cells also secrete cytokines, and there are indications that CD8⫹ T cells may also differentiate into more than one killer subtype, the diversity of CD8⫹ T-cell effector functions are clearly more restricted than those of CD4⫹ TH cells. The generation and activity of CTL cells are described in more detail in Chapter 13. Helper T Cells Can Be Divided into Distinct Subsets Tim Mosmann, Robert Coffman, and colleagues can be credited with one of the earliest experiments definitively demonstrating that helper CD4⫹ T cells were more heterogeneous in phenotype and function than originally supposed. Earlier investigations, showing that helper T cells produced a diverse array of cytokines, hinted at this possibility. However, Mosmann and Coffman definitively identified two distinct functional subgroups, TH1 and TH2, each of which produced a different set of cytokines. Furthermore, they showed that these differences were properties of distinct T-cell clones: each activated T cell expanded into a population of effector T cells that secreted a distinct array of cytokines. Specifically, these investigators developed over 50 individual T-cell clones from a mixture of T cells with different receptor specificities (i.e., a polyclonal T-cell population) that had been isolated from the spleen of an immunized mouse. At a time when the community did not have the tools to identify most cytokines directly, these researchers developed clever (and elaborate) strategies to define each clone’s cytokine secretion pattern. They showed that the cytokines secreted by each of the 50 clones fell into one of two broad categories, which they named TH1 and TH2. Because TH1 and TH2 subsets were originally identified from in vitro studies of cloned T-cell lines, some researchers doubted that they represented true in vivo subpopulations. They suggested instead that these subsets might represent different maturational stages of a single lineage. In addition, the initial failure to locate either subset in humans led some to believe that TH1, TH2, and other subsets of T helper cells were not present in this species. Further research corrected these views. In many in vivo systems, the full commitment of populations of T cells to either a TH1 or TH2 phenotype occurs late in an immune response. Hence, it was difficult to find clear TH1 or TH2 subsets in studies employing healthy human subjects, where these cells would not have developed. Indeed, TH1 and TH2 populations in T cells were ultimately isolated from humans during chronic infectious disease or chronic episodes of allergy, and studies in humans and mice definitively confirmed their lineage independence. With the benefit of new tools and technology, we now have a more detailed understanding of the panels of cytokines that each group produces. The TH1 subset secretes IL-2, IFN-␥, and Lymphotoxin-␣ (TNF-␤), and is responsible for many classic cell-mediated functions, including activation of cytotoxic T lymphocytes and macrophages. The TH2 subset secretes IL-4, IL-5, IL-6, IL-9, IL-10, and IL-13, and regulates B-cell activity and differentiation. These experiments set the stage for the discovery over the last decade that CD4⫹ T cells can adopt not just two but at least five distinct effector fates after activation. The TH1 and TH2 subpopulations have been joined definitively by the TH17 and TREG subpopulations, each of which produces a distinct cytokine profile and regulates distinct activities within the body. Yet another subpopulation, T follicular helper cells (TFH), has recently been characterized, and has achieved membership among the major helper subsets. More are bound to reveal themselves, although it will be important to determine whether each represents distinct subgroups or variants within the major subgroups. In retrospect, we probably should have anticipated the heterogeneity of helper responses, which allows an organism T-Cell Activation, Differentiation, and Memory to “tailor” a response to a particular type of pathogen. The type of effector TH cell that a naïve T cell (also called a TH0 cell) becomes depends largely on the kind of infection that occurs. For example, extracellular bacterial infections result in the differentiation of activated CD4⫹ T cells into TH2 cells, which help to activate B cells to secrete antibodies that can opsonize bacteria and neutralize the toxins they produce. On the other hand, infection by an intracellular virus or bacterium induces differentiation of CD4⫹ T cells into TH1 helpers that enhance the cytolytic activity of macrophages and CD8⫹ T cells, which can then kill infected cells. Responses to fungi stimulate the differentiation of different helper responses than responses to worms, and so on. The reality is, of course, more complex. Infections evoke the differentiation of more than one helper subtype, some of which have overlapping roles. What regulates the differentiation of each effector subset and what function each subset plays are still being actively investigated. We describe the fundamentals of our current understanding below. The Differentiation of T Helper Cell Subsets Is Regulated by Polarizing Cytokines As you know, T-cell activation requires TCR and costimulatory receptor engagement, both of which are supplied by an activated APC. It is now clear that the functional fate of activated T cells is determined by signals they receive from additional cytokines generated during the response. These cytokines (Signal 3) are referred to as polarizing cytokines because they are responsible for guiding a helper T cell toward one of several different effector fates. For example, T cells that are activated in the presence of IL-12 and IFN-␥ tend to differentiate, or polarize, to the TH1 lineage, whereas T cells that are activated in the presence of IL-4 and IL-6 polarize to the TH2 lineage. Polarizing cytokines can be generated by the stimulating APC itself, or by neighboring immune cells that have also been activated by antigen. Which cytokines are produced during an immune response depends on (1) the cell of origin (DC, macrophage, B cell, NK cell, etc.), (2) its maturation and activation status, (3) which pathogens and other inflammatory mediators it encounters, and (4) in what tissue environment it encounters that pathogen. Innate interactions therefore have a critical role in shaping adaptive responses (see Figure 5-18). Specifically, by influencing APC secretions and the surface and the microenvironmental landscape that a T cell encounters, innate immune responses directly influence the functional fate of helper T cells. Recall from Chapter 5 that APCs and other innate immune cells are activated by interaction with pathogens bearing pathogen-associated molecular patterns (PAMPs). These PAMPs bind PRRs, including, but not limited to Tolllike receptors (TLRs). PRR interactions activate dendritic cells by stimulating the up-regulation of MHC and costimulatory proteins. They also determine the type of cytokine(s) that dendritic cells and other immune cells will secrete. | CHAPTER 11 PAMPs from pathogens (or adjuvants) 371 PRR Other cellular sources of cytokines PRR APC Polarizing cytokines STATs TCR Effector cytokines Master gene regulators FIGURE 11-8 General events and factors that drive TH subset polarization. Interaction of pathogen with pattern recognition receptors (PRRs) on dendritic cells and other neighboring immune cells determines which polarizing cytokines are produced and, hence, into which T helper subset a naïve cell will differentiate. In general, polarizing cytokines that arise from dendritic cells or other neighboring cells interact with cytokine receptors and generate signals that induce transcription of unique master gene regulators. These master regulators, in turn, regulate expression of various genes, including effector cytokines, which define the function of each subset. Hence, PRR signals regulate the fate a T cell will adopt following activation (Figure 11-8). For example, double-stranded RNA, a product of many viruses, binds TLR3 receptors on dendritic cells, initiating a signaling cascade that results in production of IL-12, which directly promotes TH1 differentiation. On the other hand, worms stimulate PRRs on innate immune cells, including mast cells, which generate IL-4. IL-4 directly promotes polarization of activated T cells to the TH2 subset, which coordinates the IgE response to helminths (see Figure 11-8). In this case, the main polarizing cytokine is not made by the activating dendritic cell, but is generated by a neighboring immune cell. The pathogen interactions that give rise to the polarizing cytokines that drive T helper cell differentiation are often complex and a very active area of research. Adjuvants, which have been used for decades to enhance the efficacy of vaccines, are now understood to exert their influence on the innate immune system by regulating the expression of costimulatory ligands and cytokines by APCs, events that ultimately shape the consequences of T-cell activation. PAMPs and cytokines such as IL-12, produced by APCs themselves, are considered natural adjuvants. Dead mycobacteria, which clearly activate many PRRs, have long PA R T V 372 TABLE 11-3 | Adaptive Immunity: Effector Responses Regulation and function of T helper subtypes Polarizing cytokines Master gene regulators Effector cytokines Functions TH1 IL-12 IFN-␥ IL-18 T-Bet IFN-␥ TNF Enhances APC activity Enhances TC activation Protects against intracellular pathogens Involved in delayed type hypersensitivity, autoimmunity TH2 IL-4 GATA-3 IL-4 IL-5 IL-13 Protects against extracellular pathogens (particularly IgE responses) Involved in allergy TH17 TGF-␤ IL-6 (IL-23) ROR␥ IL-17A IL-17F IL-22 Protects against some fungal and bacterial infections Contributes to inflammation, autoimmunity TREG TGF-␤ IL-2 FoxP3 IL-10 TGF-␤ Inhibits inflammation TFH IL-6 IL-21 Bcl-6 IL-4 IL-21 B cell help in follicles and germinal centers been used as a very potent adjuvant for immune responses in mice. Very few adjuvants are approved for human vaccination, but given our new and evolving understanding of the molecules that stimulate PRRs and the consequences of that stimulation, investigators expect that we will one day be able to shape the effector response to vaccine antigens by varying the adjuvants—natural and/or synthetic—included in vaccine preparations (see Chapter 17). Effector T Helper Cell Subsets Are Distinguished by Three Properties Each helper T-cell subset is defined by an array of features, the details of which can rapidly overwhelm a new (or old) student of immunology. Understanding these specifics is an important first step to clarifying the role each subset plays in resolving infection and causing disease. Having a reference to them is also helpful when deciphering primary literature describing advances. However, some generalizations provide a useful conceptual framework for organizing some of these details. Consider the following: • Each of the major T helper cell subsets is characterized by (1) a distinct set of polarizing cytokines that induce the expression of (2) a master gene regulator that regulates expression of (3) a signature set of effector cytokines the T-cell population produces once it is fully differentiated (see Figure 11-8 and Table 11-3). • Which effector subset an activated helper cell becomes depends on the quality and quantity of signals its naïve cell precursor receives from APCs in a secondary lymphoid organ; that activity, in turn, depends on the nature of the pathogen the APC encountered at the site of infection. • Broadly speaking, TH1 and TH17 cells regulate cellmediated immunity (CD8⫹ T cells and macrophages) and TH2 and TFH cells regulate humoral immunity (B cells). However, it is important to recognize that all CD4⫹ effector T-cell subsets may have the potential to provide help to B cells. TH1 and TH17 subsets generally encourage B cells to produce antibodies that contribute to cell-mediated immunity (e.g., isotypes like IgG2a that can “arm” NK cells for cytotoxicity; see Chapter 13). TH2 cells encourage B cells to produce antibodies that mediate the clearance of extracellular pathogens (e.g., isotypes like IgE that induce the release of molecules that harm extracellular parasites). • Helper T-cell subsets often “cross-regulate” each other. The cytokines they secrete typically enhance their own differentiation and expansion and inhibit commitment to other helper T-cell lineages. This is particularly true of the TH1 and TH2 pair, as well as the TH17 and TREG pair. • Helper cell lineages may not be fixed; some subsets can assume the cytokine secretion profile of other subsets if exposed to a different set of cytokines, particularly early in the differentiation process. • The precise biological function and sites of differentiation and activity of each subset continue to be actively investigated. Much remains unknown. We start our discussion of helper cell subset characteristics with the first two subsets to be identified: TH1 and TH2 cells. They provide an illustrative example of the features that T-Cell Activation, Differentiation, and Memory Pathogens inducing cell-mediated immunity TH1-polarizing (most viruses, some factors bacteria and fungi) IL-12 PRRs Dendritic cell | CHAPTER 11 Naïve TH cell T H1 TCR MHC Class IIpeptide Signal 1 T-Bet Signal 2 GATA-3 TH2-polarizing Signal 3 TH2 TH2-polarizing factors FIGURE 11-9 Regulation of TH1 and TH2 subset differentiation. This figure depicts some of the cellular events that drive TH1 and TH2 lineage commitment in more detail. Intracellular pathogens activate a cascade of signals that polarize cells to the TH1 lineage. For example, viruses interact with PRRs (e.g., TLR-3) that induce dendritic cells to generate IL-12. This binds to receptors on naïve T cells, activating a signal transduction pathway mediated by STAT4 that induces expression of the transcription factor T-Bet. T-Bet, in turn, activates expression of effector cytokines, including IFN-␥, which define the TH1 subset’s functional capacities (and can also enhance TH1 polarization). distinguish T helper cells as well as the relationship between subsets. We follow this section with summaries of what is currently understood about the more recently characterized helper subsets: TH17, TFH, and TREG. The Differentiation and Function of TH1 and TH2 Cells The key polarizing cytokines that induce differentiation of naïve T cells into TH1 cells are IL-12, IL-18, and IFN-␥ (Figure 11-9). IL-12 is produced by dendritic cells after an encounter with pathogens via PRRs (e.g., TLR4, TLR3). It is also up-regulated in response to IFN-␥, which is generated by activated T cells and activated NK cells. IL-18, which is also produced by dendritic cells, promotes proliferation of developing TH1 cells and enhances their own production of IFN-␥. These polarizing cytokines trigger signaling pathways that up-regulate the expression of the master gene regulator T-Bet. This master transcription factor induces commitment to the TH1 lineage, inducing expression of the signature TH1 effector cytokines, including IFN-␥ and TNF. IFN-␥ is a particularly potent effector cytokine. It activates macrophages, stimulating these cells to increase microbicidal activity, up-regulate the level of class II MHC, and, as mentioned above, secrete cytokines such as IL-12, which further enhance TH1 differentiation. IFN-␥ secretion also induces antibody class switching in B cells to IgG classes (such as IgG2a in the mouse) that support phagocytosis and fixation of complement. Finally, IFN-␥ secretion promotes IFN-γ TH1-polarizing Signal 3 CD80/86 CD28 Pathogens inducing humoral immunity, particularly extracellular parasites (e.g., worms) 373 IL-4 IL-4 IL-5 IL-13 Cellular sources of IL-4 On the other hand, extracellular pathogens activate signal cascades that can polarize naïve T cells to the TH2 lineage. Parasitic worms interact with PRRs on neighboring immune cells (such as mast cells, basophils, or germinal center B cells), triggering the release of the signature TH2 polarizing cytokine IL-4. This interacts with receptors on T cells that activate STAT6, up-regulating expression of the transcriptional regulator GATA-3. GATA-3, in turn, induces expression of the TH2 effector cytokines, including IL-4, IL-5, and IL-13. [Adapted from M. L. Kapsenberg, 2003, Dendritic-cell control of pathogen-driven T-cell polarization, Nature Reviews Immunology 3:984.] the differentiation of fully cytotoxic TC cells from CD8⫹ precursors by activating the dendritic cells that engage naïve TC cells. These combined effects make the TH1 subset particularly suited to respond to viral infections and other intracellular pathogens. They also contribute to the pathological effects of TH1 cells, which are also involved in the delayed type hypersensitivity response to poison ivy (see Chapter 15). Just as differentiation to the TH1 subset is promoted by IL-12 and IFN-␥, differentiation to the TH2 subset is promoted by a defining polarizing cytokine, IL-4 (see Figure 11-9). Exposing naïve helper T cells to IL-4 at the beginning of an immune response causes them to differentiate into TH2 cells. Interestingly, TH2 development is greatly favored over TH1 development. Even in the presence of IFN-␥ and IL-12, naïve T cells will differentiate into TH2 effectors if IL-4 is present. IL-4 triggers a signaling pathway within the T cell that up-regulates the master gene regulator GATA3, which, in turn, regulates expression of TH2-specific cytokines, including IL-4, IL-5, and IL-13. Dendritic cells do not make IL-4, so from where does it come? Mast cells, basophils, and NKT cells can be induced to make IL-4 after exposure to pathogens and could influence helper T cell fate in the periphery. Germinal-center B cells and TFH cells can also produce IL-4, which could influence helper T-cell polarization in the lymph nodes and spleen. And TH2 cells themselves are an excellent source of additional IL-4 that can enhance polarization events. 374 PA R T V | Adaptive Immunity: Effector Responses Investigators, however, are still working to definitively identify the source of the IL-4 that initiates TH2 polarization in secondary lymphoid tissues. The effector cytokines produced by TH2 cells help to clear extracellular parasitic infections, including those caused by helminths. IL-4, the defining TH2 effector cytokine, acts on both B cells and eosinophils. It induces eosinophil differentiation, activation, and migration and promotes B-cell activation and class switching to IgE. These effects act synergistically because eosinophils express IgE receptors (Fc␧R) which, when cross-linked, release inflammatory mediators that are particularly good at attacking roundworms. Thus, IgE antibody can form a bridge between the worm and the eosinophil, binding to worm antigens via its variable regions and binding to Fc␧R via its constant region. IgE antibodies also mediate allergic reactions, and the role of TH2 activity in these pathological responses is described in Chapter 15. IL-5 can also induce B-cell class switching to IgG subclasses that do not activate the complement pathway (e.g., IgG1 in mice), and IL-13 functions largely overlap with IL-4. Finally, IL-4 itself can suppress the expansion of TH1-cell populations. TH1 and TH2 Cross-regulation The major effector cytokines produced by TH1 and TH2 subsets (IFN-␥ and IL-4, respectively) not only influence the overall immune response, but also influence the helper T cell subsets. First, they promote the growth (and in some cases even the polarization) of the subset that produces them; second, they inhibit the development and activity of the opposite subset, an effect known as cross-regulation. For instance, IFN-␥ (secreted by the TH1 subset) inhibits proliferation of the TH2 subset, and IL-4 (secreted by the TH2 subset) downregulates the secretion of IL-12 by APCs, thereby inhibiting TH1 differentiation. Furthermore, IL-4 enhances TH2 cell development by making TH cells less susceptible to the TH1 promoting cytokine signals (and vice versa). Similarly, these cytokines have opposing effects on target cells other than TH subsets. In mice, where the TH1 and TH2 subsets have been studied most extensively, the cytokines have distinct effects on the type of antibody made by B cells. Recall from Chapter 3 (and see Chapter 13) that the antibody isotype IgG2a enhances cell-mediated immunity by arming NKT cells, whereas the isotypes IgG1 and IgE enhance humoral immunity by their activities on extracellular pathogens. IFN-␥ secreted by the TH1 subset promotes IgG2a production by B cells but inhibits IgG1 and IgE production. On the other hand, IL-4 secreted by the TH2 subset promotes production of IgG1 and IgE and suppresses production of IgG2a. Thus, these effects on antibody production are consistent with TH1 and TH2 subsets’ overall tendencies to promote cell-mediated versus humoral immunity, respectively. IL-10 secreted by TH2 cells also inhibits (cross-regulates) TH1 cell development, but not directly. Instead, IL-10 acts on monocytes and macrophages, interfering with their ability to activate the TH1 subset by abrogating their activation, APC TCR IL-12 Stat4 T-Bet − IL-4 Stat6 − GATA-3 + IFN-γ − − IL-4 + − IL-5 + Promotes TH1 Promotes TH2 FIGURE 11-10 Cross-regulation of T helper cell subsets by transcriptional regulators. GATA-3 and T-Bet reciprocally regulate differentiation of TH1 and TH2 lineages. IL-12 promotes the expression of the TH1-defining transcription factor, T-Bet, which induces expression of TH1 effector cytokines, including IFN-␥. At the same time, T-Bet represses the expression of the TH2 defining master transcriptional regulator, GATA-3, as well as expression of the effector cytokines IL-4 and IL-5. Reciprocally, IL-4 promotes expression of GATA-3, which up-regulates the synthesis of IL-4 and IL-5, and at the same time represses the expression of T-Bet and the TH1 effector cytokine IFN␥. [Adapted from J. Rengarajan, S. Szabo, and L. Glimcher, 2000, Transcriptional regulation of Th1/Th2 polarization, Immunology Today 21:479.] specifically by (1) inhibiting expression of class II MHC molecules, (2) suppressing production of bactericidal metabolites (e.g., nitric oxide) and various inflammatory mediators (e.g., IL-1, IL-6, IL-8, GM-CSF, G-CSF, and TNF-␤), and even by (3) inducing apoptosis. The master regulators T-Bet and GATA-3 also play an important role in cross-regulation. (Figure 11-10) Specifically, the expression of T-Bet drives cells to differentiate into TH1 cells and suppresses their differentiation along the TH2 pathway. Expression of GATA-3 does the opposite, promoting the development of naïve T cells into TH2 cells while suppressing their differentiation into TH1 cells. Consequently, cytokine signals that induce one of these transcription factors sets in motion a chain of events that represses the other. TH17 Cells The discovery of the TH17 subset of T cells, which, like the TH1 subset, is involved in cell-mediated immunity, arose in part from a recognition that IL-12, one of the polarizing cytokines that induces TH1 development, was a member of a family of cytokines that shared a subunit (p40) with IL-23. The p40 knockout mouse was a favorite model for studying the importance of TH1 cells because, in the absence of IL-12, it failed to generate TH1 cells. However, once it was understood T-Cell Activation, Differentiation, and Memory | CHAPTER 11 375 11-11 OVERVIEW FIGURE T Helper Subset Differentiation Naïve CD4+ T cell Polarizing cytokines Master transcriptional regulator Effector cytokines Effector functions IL-2, TGF-β IL-1, IL-6, IL-23, TGF-β IL-4 IL-6, IL-21 Bcl-6 IL-12, IFN-γ, IL-18 T-Bet FOXP3 RORγt GATA3 Induced TReg cell TH17 cell TH2 cell TFH cell TH1 cell IL-10, TGF-β IL-17A, IL-17F, IL-22 IL-4, IL-5, IL-13 IL-4, IL-21 IFN-γ, TNF Regulation, suppression of immune and inflammatory responses Inflammation Allergic and anti-helminth responses B cell help in germinal centers Cell-mediated immunity, macrophage activation, inflammation This figure synthesizes current information about the distinguishing features of T helper subset differentiation and activity. Polarizing cytokines, master transcriptional regulators, effector cytokines, and broad functions in health and disease are depicted for each of the major helper subsets. Neither cross-regulation nor the potential that p40 was also required for IL-23 production, investigators quickly realized that the results from these mice were no longer unambiguous. In fact, it became clear that these mice also failed to produce a T-cell subset that required IL-23. Some of the functions originally attributed to IL-12-induced TH1 cells were actually carried out by an IL-23-induced T helper cell subpopulation now referred to as TH17 cells. TH17 cells are generated when naïve T cells are activated in the presence of IL-6 and TGF-␤, the key polarizing cytokine for iTREG differentiation (Overview Figure 11-11). IL-23 also plays a role in finalizing the commitment to the TH17 fate and is induced in APCs by interactions with PAMPs including fungal wall components, with TLR2 and the CLR Dectin-1. Like TH1 and TH2 differentiation, TH17 cell differentiation is also controlled by a master transcriptional regulator whose expression is induced by polarizing cytokines. In this case the master regulator is the orphan steroid receptor ROR␥t, plasticity in differentiation among subsets is depicted, but both are described in the text. [Adapted from S. L. Swain, K. K. McKinstry, and T. M. Strutt, Expanding roles for CD41 T cells in immunity to viruses, Nature Reviews. Immunology 12:136–148.] which also plays a role in T-cell development. ROR␥t knockout mice have reduced severity of experimental autoimmune encephalitis (EAE, a mouse model of multiple sclerosis) apparently because of the reduction in TH17 cells. TH17 cells are so named because they produce IL-17A, a cytokine associated with chronic inflammatory and autoimmune responses, including those that result in inflammatory bowel disease, arthritis, and multiple sclerosis. TH17 cells are the dominant inflammatory cell type associated with chronic autoimmune disorders. They also produce IL-17F and IL-22, cytokines associated with tissue inflammation. We have only begun to understand the true physiological function of TH17 cells, which in healthy humans have been found at epithelial surfaces (e.g., lung and gut) and may play a role in warding off fungal and extracellular bacterial infections (see Clinical Focus Box 11-4). However, a full appreciation of their biological role awaits further investigations. 376 PA R T V | Adaptive Immunity: Effector Responses CLINICAL FOCUS What a Disease Reveals about the Physiological Role of TH17 Cells Experiments that reveal the inner workings of normal healthy immune cells have offered invaluable insights into what can and does go wrong. However, at times we need “experiments of nature”— the diseases themselves—to clarify how the immune system works in healthy individuals. Job syndrome, a rare disease in which patients suffer from defects in bone, teeth, and immune function, is helping us solve the mystery of the physiological function of TH17 cells. People with Job syndrome suffer from recurrent infections, typically of the lung and skin. The painful abscesses that are often a feature of the disease and the trials endured by its patients explain its name, which comes from a biblical story of a man (Job) who is subject to horrific hardship as a test of his piety. Another cardinal feature of the disease, elevated IgE levels in serum, is the basis for its more formal name, Hyper IgE Syndrome (HIES). HIES comes in two major forms: Job syndrome, the most common, is also referred to as Type 1 HIES. Patients with Type 2 HIES, which we will not discuss here, do not have trouble with bone or dental development. The abundance of IgE originally suggested to some investigators who were savvy about the roles of helper T-cell subsets that Type 1 HIES symptoms may be caused by an imbalance between TH1 and TH2 responses. Initial work did not reveal a difference in these activities, but as the diversity of T helper subsets was revealed, investigators pursued the possibility of a helper imbalance more vigorously. A number of groups from Japan and America independently discovered that lymphocytes from HIES patients were unable to respond to select cytokines, including IL-6, IL-10, and IL-23. An analysis of genes that could explain this signaling failure revealed a dominant negative mutation in STAT3, a key cytokine signaling molecule (see Chapter 4). These investigators knew from the literature that the absence of signaling through these cytokines suggested a specific problem with polarization to the TH17 helper subset. Indeed, circulating TH17 cells were absent in HIES patients with STAT3 mutations. The investigators then directly tested their hypothesis that this absence was due to a failure of CD4⫹ T cells to polarize normally. Specifically, they removed T cells from blood, and exposed them to conditions that would ordinarily polarize them to the TH17 lineage: T-cell receptor stimulation (Signal 1) in the presence of costimulatory signals (CD28, Signal 2) and cytokines known to drive human TH17 differentiation (IL-1, IL-6, TGF-␤, and IL-23, (Induced) TREG Cells Another major CD4⫹ T-cell subset negatively regulates T-cell responses and plays a critically important role in peripheral tolerance by limiting autoimmune T-cell activity. This subset of T cells, designated induced TREG (iTREG) cells, is similar in function to the natural TREG cells (nTREGs) that originate from the thymus (see Chapter 9). Induced TREG cells, however, do not arise in the thymus, but from naïve T cells that are activated in secondary lymphoid tissue in the presence of TGF-␤ (see Overview Figure 11-11). TGF-␤ induces expression of FoxP3, the master transcriptional regulator responsible for iTREG commitment. The iTREG cells secrete the effector cytokines IL-10 and TGF␤, which down-regulate inflammation via their inhibitory effects on APCs, and can also exert their suppressive func- Signal 3) (Figure 1a). They stained the cells for intracellular expression of TH17’s signature cytokine, IL-17, and performed flow cytometry (Figure 1b; also see Chapter 20). Although CD4⫹ T cells from HIES patients were able to develop into other helper lineages (and, as you can see from the flow cytometry contour plots, were also able to make IFN-␥, indicating they could become TH1 cells), they could not be induced to secrete IL-17. Specifically, whereas 18.3% of T cells from healthy patients that were subject to these conditions stained with antibodies against IL-17, 0.05% (essentially none) of the T cells from HIES patients stained with the same antibodies. These striking observations suggest that the recurrent infections HIES patients experience are caused at least in part by the absence of TH17 cells. Reciprocally, they indicate that TH17 cells play an important role in controlling the type of infection that afflicts these patients, including Staphylococcus aureus skin infections and pneumonia. The critical role of TH17 in controlling bacterial and fungal infections at epithelial surfaces has been supported by studies in mice, too. J. D. Milner et al. 2008. Impaired TH17 cell differentiation in subjects with autosomal dominant Hyper-IgE syndrome. Nature 452:773–776. tion by interacting directly with T cells. The depletion of iTREG cells in otherwise healthy animals leads to multiple autoimmune outbreaks, revealing that even healthy organisms are continually warding off autoimmune responses. Recent data also indicate that iTREG cells are critically important for maintaining a mother’s tolerance to her fetus. TH17 and TREG Cross-Regulation Just as TH1 and TH2 cells reciprocally regulate each other, TREG and TH17 cells also cross-regulate each other. TGF-␤ induces TREG differentiation; however, when accompanied by IL-6, TGF-␤ induces TH17 differentiation. Specifically TGF-␤ appears to up-regulate both FoxP3 and ROR␥ (which control TREG and TH17 differentiation, respectively). In combination with signals generated by IL-6, T-Cell Activation, Differentiation, and Memory | CHAPTER 11 377 BOX 11-4 IL-17? TCR and CD28 engagement +/– polarizing cytokines (IL-1/IL-23) 12 days Naïve CD4+ T cells TH17 cells? (a) (b) FIGURE 1 CD4ⴙ T cells from Hyper-IgE patients do not differentiate into TH17 cells. (a) Condi- tions used to polarize CD4⫹ T cells to the TH17 lineage in vitro. The ability of the cells to make IL-17 (a feature of TH17 cells) or IFN-␥ (a property of TH1, not TH17 cells) after they were polarized was assessed by staining for intracellular cytokines. (b) Flow cytometric analysis of intracellular staining. The boxes outlined in red indicate the quadrant where IL-17⫹ (TH17) cells would appear. [Irvine, D.J. et al. 2002. Nature 419: 845–849. Reprinted by permission from Macmillan Publishers] signals generated by TGF-␤ inhibit FoxP3 expression, letting ROR␥ dominate and induce TH17 development. The TH17 versus iTREG relationship may be very adaptive. At rest, a healthy organism may favor the development of an anti-inflammatory iTREG population, which would be reinforced by the iTREG cell’s own production of TGF-␤. Inflammation, however, would induce the generation of acute response proteins, including IL-6. In the presence of IL-6, TGF-␤ activity would shift development of T cells away from iTREGs toward the pro-inflammatory TH17, so a proper defense could be mounted. TFH Cells Follicular helper T (TFH) cells are a very recent addition to the helper T-cell subset family. Whether they represent an independent lineage or a developmental option for other helper lineages remains controversial. Like TH2 cells, TFH cells play a central role in mediating B-cell help and are found in B-cell follicles and germinal centers. However, the effector cytokines secreted by follicular helper T cells differ partially from those secreted by TH2 cells. Cytokines that polarize activated T cells toward the TFH lineage include IL-6 and IL-21. These polarizing cytokines induce the expression of Bcl-6, a transcriptional repressor that is thought to be TFH’s master transcriptional regulator (see Overview Figure 11-11). Cross-regulation is also a feature of TFH function: Bcl-6 expression inhibits T-bet, GATA3, and ROR␥t expression, thus inhibiting TH1, TH2, and TH17 differentiation, respectively, while inducing TFH polarization. 378 PA R T V | Adaptive Immunity: Effector Responses Although both TFH and TH2 cells secrete IL-4, TFH cells are best characterized by their secretion of IL-21, which induces B-cell differentiation. Interestingly, they can also produce IFN-␥ (the defining TH1 cytokine). How TFH and TH2 cells divide responsibilities for inducing B-cell antibody production is still an open question. Other Potential Helper T-Cell Subsets Other T-cell subsets with distinct polarizing requirements and unique cytokine secretion profiles have been identified (e.g., TH9 cells, which secrete IL-9 and IL-10). However, because these subpopulations secrete cytokines that are also produced by TH1, TH2, TH17, or iTREG cells, some speculate that these cell types do not represent distinct subclasses but rather are developmental or functional variants of one of the major subpopulations. This perspective has, indeed, been applied to the follicular helper T-cell (TFH) subset, which also expresses cytokines shared by several other subtypes. However, this subset has a distinct gene signature and a distinct master regulator (bcl-6), so most now consider it a bona fide independent lineage. Clearly, our understanding of the developmental relationship among effector subtypes will continue to evolve. Helper T Cells May Not Be Irrevocably Committed to a Lineage Investigations now suggest that the relationship among TH cell subpopulations may be more plastic than previously suspected. At early stages in differentiation, at least, helper cells may be able to shift their commitment and produce another subset’s signature cytokine(s). For example, when exposed to IL-12, young TH2 cells can be induced to express the signature TH1 cell cytokine, IFN-␥. Young TH1 cells can also be induced to express the signature TH2 cytokine, IL-4, under TH2 polarizing conditions. Interestingly, TH1 and TH2 cells do not seem able to adopt TH17 or iTREG characteristics. On the other hand, TH17 and iTREG cells are more flexible and can adopt the cytokine expression profiles of other subsets, including TH1 and TH2. The TH17 subset may be the most unstable or “plastic” lineage and can be induced to express IFN-␥ and IL-4, depending on environmental input. Some iTREG cells can also be induced to express IFN-␥, and some can be redirected toward a TH17 phenotype if exposed to IL-6 and TGF-␤. This fluidity among subsets makes it difficult to definitively establish the independence of helper cell lineages. In fact, some of the emerging subgroups may be variants of TH1, TH2, TH17, and iTREG subsets that have been exposed to other polarizing environments. Helper T-Cell Subsets Play Critical Roles in Immune Health and Disease Studies in both mice and humans show that the balance of activity among T-cell subsets can significantly influence the outcome of the immune response. A now classic illustration of the influence of T-cell subset balance on disease outcome is provided by leprosy, which is caused by Mycobacterium leprae, an intracellular pathogen that can survive within the phagosomes of macrophages. Leprosy is not a single clinical entity; rather, the disease presents as a spectrum of clinical responses, with two major forms of disease, tuberculoid and lepromatous, at each end of the spectrum. In tuberculoid leprosy, cell-mediated immune responses destroy most of the mycobacteria. Although skin and peripheral nerves are damaged in tuberculoid leprosy, it progresses slowly and patients usually survive. In contrast, lepromatous leprosy is characterized by a humoral response; cell-mediated immunity is depressed. The humoral response sometimes results in markedly high levels of immunoglobulin (hypergammaglobulinemia). This response is not as effective in inhibiting disease, and mycobacteria are widely disseminated in macrophages, often reaching numbers as high as 1010 per gram of tissue. Lepromatous leprosy progresses into disseminated infection of the bone and cartilage with extensive nerve and tissue damage. The development of lepromatous or tuberculoid leprosy depends in part on the balance between TH1 and TH2 responses (Figure 11-12). In tuberculoid leprosy, the immune response is characterized by a TH1-type response and high circulating levels of IL-2, IFN-␥, and Lymphotoxin-␣ (LT-␣). In lepromatous leprosy, there is a TH2-type immune response, with high circulating levels of IL-4 and IL-5 (and IL-10, which can also be made by TH2 cells). This cytokine profile explains the diminished cell-mediated immunity and increased production of serum antibody in lepromatous leprosy. Presumably each of these patients was exposed to the same pathogen. Why did some develop an effective TH1 response while others did not? Studies suggest that genetic differences among human hosts play a role. For example, differences in susceptibility may correlate with individual differences in the expression of PRRs (TLR1 and TLR2) expressed TH1 activity Tuberculoid Lepromatous TH2 activity Tuberculoid Lepromatous IL-2 IL-4 IFN-␥ IL-5 LT-␣ IL-10 FIGURE 11-12 Correlation between type of leprosy and relative TH1 or TH2 activity. Messenger RNA isolated from lesions from tuberculoid and lepromatous leprosy patients was analyzed by Southern blotting using the cytokine probes indicated. Cytokines characteristic of TH1 cells (IFN-␥ and TNF-␤, for instance) predominate in the tuberculoid patients, whereas cytokines characteristic of TH2 cells (IL-4) predominate in the lepromatous patients. [From P. A. Sieling and R. L. Modlin, 1994, Cytokine patterns at the site of mycobacterial infection, Immunobiology 191:378.] T-Cell Activation, Differentiation, and Memory by innate cells. This makes sense given that interactions between pathogen and innate immune cells determine the cytokine environment that influences the outcome of T-cell polarization. Differences in TLR expression or activity could alter the quality or quantity of cytokines produced. Progression of HIV infection to AIDS may also be influenced by T-cell subset balance. Early in the disease, TH1 activity is high, but as AIDS progresses, some have suggested that a shift may occur from a TH1-like to a TH2-like response, which is less effective at controlling viral infection. In addition, some pathogens may “purposely” influence the activity of the TH subsets. The Epstein-Barr virus, for example, produces a homolog (mimic) of human IL-10 called viral IL-10 (vIL-10). Like cellular IL-10, vIL-10 tends to suppress TH1 activity by inhibiting the polarizing activation of macrophages. Some researchers have speculated that vIL-10 may reduce the cell-mediated response to the Epstein-Barr virus, thus conferring a survival advantage on the virus. TH17 cells first received attention because of their association with chronic autoimmune disease. Mice that were unable to make IL-23, a cytokine that contributes to TH17 polarization, were remarkably resistant to autoimmunity. TH17 cells and their defining effector cytokine, IL-17, are often found in inflamed tissue associated with rheumatoid arthritis, inflammatory bowel disease, multiple sclerosis, psoriasis, and asthma. However, the role TH17 cells played in protecting organisms from infection was not immediately obvious. Studies of individuals with an autosomal dominant form of a disease known as Hyper-IgE syndrome or Job syndrome confirmed indications from mice that TH17 cells were important in controlling infections by extracellular bacteria and fungi (see Clinical Focus Box 11-4). These disorders and those described in Chapters 15 and 16 are just some examples of the influence of helper T-cell subsets on disease development. It is important to recognize that our current perspectives of the roles of helper subsets in disease and health are still simplistic. Our developing appreciation of the complexity of the interplay among subsets will improve and add more subtlety to our explanations in the future. T-Cell Memory T-cell activation results in a proliferative burst, effector cell generation, and then a dramatic contraction of cell number. At least 90% of effector cells die by apoptosis after pathogen is cleared, leaving behind an all-important population of antigen-specific memory T cells. Memory T cells are generally long-lived and quiescent, but respond with heightened reactivity to a subsequent challenge with the same antigen. This secondary immune response is both faster and more robust, and hence more effective than a primary response. Until recently, memory cells were difficult to distinguish from effector T cells and naïve T cells by phenotype, and for some time they were defined best by function. Like naïve T cells, most memory T cells are resting in the G0 stage of the cell cycle, but they appear to have less stringent requirements | CHAPTER 11 379 for activation than naïve T cells. For example, naïve T cells are activated almost exclusively by dendritic cells, whereas memory T cells can be activated by macrophages, dendritic cells, or B cells. Memory cells express different patterns of surface adhesion molecules and costimulatory receptors that allow them to interact effectively with a broader spectrum of APCs. They also appear to be more sensitive to stimulation and respond more quickly. This may, in part, be due to their ability to regulate gene expression more readily because of differences in epigenetic organization that occurred during their formation. Finally, memory cells display recirculation patterns that differ from those of naïve or effector T cells. Some stay for long periods of time in the lymph node and other secondary lymphoid organs, some travel to and remain in tertiary immune tissues where the original infection occurred, anticipating the possibility of another infection with the antigen to which they are specific. As our ability to identify different cell surface proteins improved, so has our ability to identify and distinguish memory cell subpopulations. We now broadly distinguish two subsets of memory T cells, central memory T cells (TCM) and effector memory T cells (TEM), on the basis of their location, their patterns of expression of surface markers, and, to some extent, their function. Recent work also reveals a great deal of diversity within these subsets, whose relationship is still being clarified. We describe some useful generalizations below, and close with the many questions that remain. Naïve, Effector, and Memory T Cells Display Broad Differences in Surface Protein Expression Three surfaces markers have been used to broadly distinguish naïve, effector, and memory T cells: CD44, which increases in response to activation signals; CD62L, an adhesion protein; and CCR7, a chemokine receptor. Both CD62L and CCR7 are involved in homing to secondary lymphoid organs (Table 11-4). Naïve T cells express low levels of CD44, reflecting their unactivated state, and high levels of the adhesion molecule CD62L, directing them to the lymph node or spleen. In contrast, effector helper and cytotoxic T cells have the reciprocal phenotype. They express high levels of CD44, indicating that they have received TCR signals, and low levels of CD62L, which prevents them from recirculating to secondary lymphoid tissue, allowing them to thoroughly probe sites of infection in the periphery. Both types of memory T cells also tend to express CD44, indicating that they are antigen experienced (i.e., have received signals through their TCR). Like naïve T cells, central memory cells (TCM) express CD62L and the chemokine receptor, CCR7, consistent with their residence in secondary lymphoid organs. Effector memory cells (TEM), which are found in a variety of tissues, can express varying levels of CD62L depending on their locale; however, they do not express CCR7, reflecting their travels through and residence in nonlymphoid tissues. Other markers have been used to distinguish subtypes of memory cells, but these still provide a useful starting point for gauging the status of a T cell. 380 PA R T V TABLE 11-4 | Adaptive Immunity: Effector Responses Surface proteins that are used to distinguish naïve, effector, and memory T cells Cell type CD44 CD62L CCR7 low ⫹ ⫹ Effector T cell ⫹ low ⫺ Effector memory T cell ⫹ variable ⫺ Central memory T cell ⫹ ⫹ ⫹ Naïve T cell TCM and TEM Are Distinguished by Their Locale and Commitment to Effector Function A small proportion (⬍10%) of the progeny of a naïve cell that has proliferated robustly in response to antigen differentiates into TCM and TEM cells. In general, these two subsets are distinguished by where they reside as well as their level of commitment to a specific effector cell fate. In general, TCM cells reside in and travel between secondary lymphoid tissues. They live longer and have the capacity to undergo more divisions than their TEM counterparts. When they reencounter their cognate pathogen in secondary lymphoid tissue, they are rapidly activated and have the capacity to differentiate into a variety of effector T-cell subtypes, depending on the cytokine environment. On the other hand, TEM cells travel to and between tertiary tissues (including skin, lung, liver, and intestine). They are arguably better situated to contribute to the first line of defense against reinfection because they have already committed to an effector lineage during the primary response and exhibit their effector functions quite rapidly after reactivation by their cognate pathogen. It is important to note that some of these generalizations may not hold up to scientific scrutiny. For instance, some investigators contest the definitive distinction between TCM and TEM cells and emphasize the diversity and continuum of variations among these subtypes. Hence, memory cell biology is one of the most active fields of investigation, one that is critical to our ability to develop the best vaccines. How and When Do Memory Cells Arise? Current work suggests that memory cells arise very early in the course of an immune response (e.g., within 3 days), but their cell of origin remains controversial. Some investigations suggest that memory cells arise as soon as naïve T cells are activated. Others suggest that memory cells arise from more fully differentiated naïve T cells. Still others raise the intriguing possibility that naïve T-cell activation generates a “memory stem cell” that is self-renewing and gives rise to memory effector cell populations. These models are not mutually exclusive, and it is possible that memory cells can arise at several different stages of T-cell activation throughout a primary response. The relationship between TCM and TEM cells is also debated. They may originate independently from naïve and effector cells, respectively, or may give rise to each other. Studies suggest, in fact, that TCM cells arise from TEM cells, and one possible model of relationships is shown in Figure 11-13. Here, investigators speculate that TCM cells arise prior to TEM cells, from cells at an earlier stage of differentiation into effector (helper or cytotoxic) T cells. TEM cells arise late, and also may develop from fully differentiated effector cells. The model also suggests that effector cells can replenish central memory cells. It should be stressed, however, that several other models have also been advanced. For instance, recent work suggests interactions experienced by effector cells determines their TCM versus TEM fate. Effector cells that interact with B cells may preferentially develop into central and not effector memory T cells. New models may also need to incorporate intriguing recent observations, including the possibilities that (1) memory Proliferation and differentiation Effector T cell Naïve T cell ? Central memory T cell FIGURE 11-13 One possible model for the development of central and effector memory T cells. This model, only one of several that have been advanced, suggests that central memory cells arise early after naïve T-cell activation, perhaps from the first divisions. Effector memory T cells may arise later, after the progeny have divided more and have assumed at least some ? Effector memory T cell effector cell features. The model also includes the possibilities that (1) some effector memory T cells arise from fully differentiated effector T cells and (2) effector memory T cells can develop into central memory T cells. [Adapted from D. Gray, 2002, A role for antigen in the maintenance of immunological memory, Nature Reviews Immunology 2:60.] T-Cell Activation, Differentiation, and Memory cells arise from the asymmetric cell division of activated T cells, where one daughter cell becomes an effector cell, and another contributes to the memory pool, and (2) that T-cell activation generates a self-renewing memory stem cell population that provides a long-term source of memory T cells. What Signals Induce Memory Cell Commitment? Most investigators agree that T-cell help is critical to generating long-lasting memory. For instance, CD8⫹ T cells can be activated in the absence of CD4⫹ T-cell help, but this “helpless” activation event does not yield long-lived memory CD8⫹ T cells. The relative importance of other variables in driving memory development is still under investigation. Although intensity of T-cell receptor engagement was thought to be a factor in memory cell commitment, recent data suggest that even low-affinity interactions can generate memory T cells. All studies, however, appear consistent with the recognition that the more proliferation a response inspires, the better the memory pool. Do Memory Cells Reflect the Heterogeneity of Effector Cells Generated during a Primary Response? We have seen that naïve T cells differentiate into a wide variety of effector T-cell subpopulations, largely determined by the cytokine signals they receive during activation. Studies indicate that the memory cell response is also very diverse, in terms of both the T-cell receptor specificities and the array | CHAPTER 11 381 of cytokines produced. However, the cellular origin of this diversity is still under investigation. Specifically, does this diverse memory response strictly reflect the functional effector diversity generated during the primary response? Or does it develop anew from central memory T cells responding to different environmental cues during rechallenge? The answer is likely to be “Both,” but investigations continue. Are There Differences between CD4⫹ and CD8⫹ Memory T Cells? The simple answer is “Maybe.” Memory CD8⫹ T cells are clearly more prevalent than memory CD4⫹ T cells. This is partly because CD8⫹ T cells proliferate more robustly and therefore generate proportionately more memory T cells. It may also be due to differences in the life span of memory T cells: CD4⫹ memory T cells may not be as long-lived as CD8⫹ memory T cells. How Are Memory Cells Maintained over Many Years? Whether memory cells can persist for years in the absence of antigen remains controversial, although evidence seems to favor the possibility that they do. Regardless, it does seem that memory persistence depends on the input of cytokines that induce occasional divisions, a process known as homeostatic proliferation, which maintains the pool size by balancing apoptotic events with cell division. Both IL-7 and IL-15 appear important in enhancing homeostatic proliferation, but CD4⫹ and CD8⫹ memory T-cell requirements may differ. S U M M A R Y ■ ■ ■ ■ T-cell activation is the central event in the initiation of the adaptive immune responses. It results from the interaction in a secondary lymphoid tissue between a naïve T cell and an APC, specifically a dendritic cell. Activation of naïve T cells leads to the differentiation of effector cells, which regulate the response to pathogen, and of long-lived memory T cells, which coordinate the stronger and quicker response to future infections by the same pathogen. Three distinct signals are required to induce naïve T-cell activation, proliferation, and differentiation. Signal 1 is generated by the interaction of the TCR-CD3 complex with an MHCpeptide complex on a dendritic cell. Signal 2 is a costimulatory signal provided by the interaction between molecules of the B7 family expressed by APC with the positive costimulatory molecules CD28 or ICOS expressed by T cells. Signal 3 is provided by soluble cytokines and plays a key role in determining the type of effector cell that a T cell becomes. In the absence of a costimulatory signal (Signal 2), T-cell receptor engagement results in T-cell inactivity or clonal anergy. ■ ■ ■ ■ CD8⫹ T cells, which recognize MHC class I-peptide complexes that are expressed by virtually all cells in the body, become cytotoxic (TC) cells with the capacity to kill infected cells. CD4⫹ T cells, which recognize MHC class II-peptide complexes that are expressed by professional APCs, become helper (TH) cells, secreting cytokines that regulate (positively and negatively) cells that clear infection, including B cells, macrophages, and other T cells. CD4⫹ T cells differentiate into at least five main subpopulations of effector cells: TH1, TH2, TH17, iTREG, and TFH. Each subpopulation is characterized by (1) a unique set of polarizing cytokines that initiate differentiation, (2) a unique master transcriptional regulator that regulates the production of helper-cell-specific genes, and (3) a distinct set of effector cytokines that they secrete to regulate the immune response. TH1 and TH17 cells generally enhance cell-mediated immunity and inflammatory responses. TH2 and TFH cells enhance humoral immunity and antibody production, and induced TREG cells inhibit T-cell responses. 382 ■ ■ ■ PA R T V | Adaptive Immunity: Effector Responses Helper T-cell subsets have also been associated with disease and play a role in the development of autoimmunity and allergy. Memory T cells, which are more easily activated than naïve cells, are responsible for secondary responses. The generation of memory B cells as well as CD4⫹ and CD8⫹ memory T cells requires T-cell help. ■ Two types of memory T cells have been described. Central memory (TCM) cells are longer lived, reside in secondary lymphoid tissues, and can differentiate into several different effector T cells. Effector memory (TEM) cells populate the sites of infection (tertiary tissues) and immediately reexpress their original effector function after reexposure to antigen. R E F E R E N C E S Ahmed, R., M. Bevan, S. Reiner, and D. Fearon. 2009. The precursors of memory: Models and controversies. Nature Reviews. Immunology 9:662–668. Miossec, P., T. Korn, and V.K. Kuchroo. 2009. Interleukin 17 and Type 17 helper T cells. New England Journal of Medicine 361:888–898. Gattinoni, L., et al. 2011. A human memory T cell subset with stem cell-like properties. Nature Medicine 17:1290–1297. Mosmann, T., H. Cherwinski, M. Bond, M. Giedlin, and R. Coffman. 1986. Two types of murine helper T cell clone. I. Definition according to profiles of lymphokine activities and secreted proteins. Journal of Immunology 136:2348– 2357. Hoyer, K. K., W. F. Kuswanto, E. Gallo, and A. K. Abbas. 2009. Distinct roles of helper T-cell subsets in systemic autoimmune disease. Blood 113:389–395. Jameson, S., and D. Masopust. 2009. Diversity in T cell memory: An embarrassment of riches. Immunity 31:859–871. Jelley-Gibbs, D., T. Strutt, K. McKinstry, and S. Swain. Influencing the fates of CD4⫹ T cells on the path to memory: Lessons from influenza. Immunology and Cell Biology 86:343–352. Kaiko, G. E., J. C. Horvat, K. W. Beagley, and P.M. Hansbro. 2007. Immunological decision making: How does the immune system decide to mount a helper T-cell response? Immunology 123: 326–338. Kapsenberg, M. L. 2003. Dendritic cell control of pathogen-driven T-cell polarization. Nature Reviews Immunology 3: 984–993. Kassiotis, G., and A. O’Garra. 2009. Establishing the follicular helper identity. Immunity 31:450–452. Murphy, K., C. Nelson, and J. Sed. 2006. Balancing co-stimulation and inhibition with BTLA and HVEM. Nature Reviews Immunology 6:671–681. Palmer, M. T., and C. T. Weaver. 2010. Autoimmunity: Increasing suspects in the CD4⫹ T cell lineup. Nature Immunology 11:36–40. Pepper, M., and M. K. Jenkins. 2011. Origins of CD4⫹ effector and memory T cells. Nature Immunology 12:467–471. Readinger, J., K. Mueller, A. Venegas, R. Horai, and P. Schwartzberg. 2009. Tec kinases regulate T-lymphocyte development and function: New insights into the roles of Itk and Rlk/Txk. Immunological Reviews 228:93–114. Reiner, S. 2008. Inducing the T cell fates required for immunity. Immunological Research 42:160–165. Khoury, S., and M. Sayegh. 2004. The roles of the new negative T cell costimulatory pathways in regulating autoimmunity. Immunity 20:529–538. Riley, J. 2009. PD-1 signaling in primary T cells. Immunology Reviews 229:114–125. King, C. 2009. New insights into the differentiation and function of T follicular helper cells. Nature Reviews. Immunology 9:757–766. Rudd, C., A. Taylor, and H. Schneider. 2009. CD28 and CTLA-4 coreceptor expression and signal transduction. Immunological Reviews 229:12–26. Korn, T., E. Bettelli, M. Oukka, and V. Kuchroo. 2009. IL-17 and TH17 Cells. Annual Review of Immunology 27:485–517. Lefrancois, L., and D. Masopust. 2009. The road not taken: Memory T cell fate “decisions.” Nature Immunology 10:369–370. Linsley, P., and S. Nadler. 2009. The clinical utility of inhibiting CD28-mediated costimulation. Immunological Reviews 229:307–321. Malissen, B. 2009. Revisiting the follicular helper T cell paradigm. Nature Immunology 10:371–372. Mazzoni, A., and D. Segal. 2004. Controlling the Toll road to dendritic cell polarization. Journal of Leukocyte Biology 75:721–730. McGhee, J. 2005. The world of TH1/TH2 subsets: First proof. Journal of Immunology 175:3–4. Sallusto, F., and A. Lanzavecchia. 2009. Heterogeneity of CD4⫹ memory T cells: Functional modules for tailored immunity. European Journal of Immunology 39:2076–2082. Sharpe, A. 2009. Mechanisms of costimulation. Immunological Reviews 229:5–11. Smith-Garvin, J., G. Koretzky, and M. Jordan. 2009. T cell activation. Annual Review of Immunology 27:591–619. Thomas, R. 2004. Signal 3 and its role in autoimmunity. Arthritis Research & Therapy 6:26–27. Thompson, C., et al. 1989. CD28 activation pathway regulates the production of multiple T-cell-derived lymphokines/ cytokines. Proceedings of the National Academy of Sciences of the United States of America 86:1333–1337. T-Cell Activation, Differentiation, and Memory Wang, S., and L. Chen. 2004. T lymphocyte co-signaling pathways of the B7-CD28 family. Cellular & Molecular Immunology 1:37–42. Weaver, C., and R. Hatton. 2009. Interplay between the TH17 and TREG cell lineages: A (co)evolutionary perspective. Nature Reviews. Immunology 9:883–889. Yu, D., et al. 2009. The transcriptional repressor Bcl-6 directs T follicular helper cell lineage commitment. Immunity 31:457–468. Zhou, L., M. Chong, and D. Littman. 2009. Plasticity of CD4⫹ T cell lineage differentiation. Immunity 30:646–655. Zhu, J., and W. Paul. 2010. Heterogeneity and plasticity of T helper cells. Cell Research 20:4–12. Useful Web Sites The following are examples of well-organized Web sites developed by undergraduate students, graduate students, and teachers of immunology who have done an excellent job S T U D Y | CHAPTER 11 383 of simplifying a complex topic: helper T-cell subset differentiation. The Web sites may not be continually updated, so, as always with Internet sources, double-check the date and the accuracy of the information. http://wenliang.my web.uga.edu/mystudy/ immunology/ScienceOfImmunology/Biological process(immuneresponses).html These Web notes from a former graduate student at the University of Georgia provide a rudimentary, but accurate and accessible, description of the T-cell subsets. http://microbewiki.kenyon.edu/index.php/Host_ Dependency_of_Mycobacterium_leprae You will find here a posting from MicrobeWiki, “a student-edited microbiology resource” originating from Kenyon College. http://users.rcn.com/jk imball.ma.ultranet/ B i o l o g y Pa g e s / T / T H 1 _ T H 2 . h t m l # Ty p e s _ o f _ Helper_T_Cells This selection is from immunologist and teacher John Kimball’s online version of his textbook. Q U E S T I O N S CLINICAL AND EXPERIMENTAL FOCUS QUESTION Multiple sclerosis is an autoimmune disease in which TH cells participate in the destruction of the protective myelin sheath around neurons in the central nervous system. Each person with this disease has different symptoms, depending on which neurons are affected, but the disease can be very disabling. Recent work in a mouse model of this disease suggests that transplantation of cell precursors of neurons may be a good therapy. Although these immature cells may work because they can develop into neuronal cells that replace the lost myelin sheath, some investigators realized that they play another, perhaps even more important role. These scientists showed TH17 TH1 0.6 2.9 TREG 1.2 36.9 Control 22.1 that the neuronal cell precursors secrete a cytokine called Leukemia Inhibiting Factor (LIF). In fact, the administration of this factor, alone, ameliorated symptoms. These investigators were curious to know if this cytokine had an effect on T-cell activity. They added LIF to cultures of (normal) T cells that were being stimulated under different polarizing conditions (i.e., TCR and CD28 engagement in the presence of cytokines that drive differentiation to distinct T helper subsets). They stained T cells for cytokine production and analyzed the results by flow cytometry. The data below show their results from normal mouse T cells polarized to the lineage indicated in the absence (top, control) or presence (bottom) of LIF. 3.4 0.7 2.7 0.6 37.7 5.2 IFN-γ 29.1 FoxP3 IL-17 LIF 11.2 27.8 CD4 PA R T V 384 | Adaptive Immunity: Effector Responses Which T helper lineage(s) is(are) most affected by the addition of LIF? Explain your answer. Why might these results explain the beneficial effect of LIF on the disease? Knowing what you know now about the molecular events that influence T-cell differentiation, speculate on the molecular basis for the activity of LIF. 1. Which of the following conditions would lead to T-cell anergy? 5. The following sentences are all false. Identify the error(s) and correct. a. Macrophages activate naïve T cells better than dendritic cells. b. ICOS enhances T-cell activation and is called a negative coreceptor. c. Virtually all cells in the body express costimulatory ligands. d. CD28 is the only costimulatory receptor that binds to B7 family members. a. A naïve T-cell interaction with a dendritic cell in the presence of CTLA-4 Ig. b. A naïve T cell stimulated with antibodies that bind both the TCR and CD28. c. A naïve T cell stimulated with antibodies that bind only the TCR. d. A naïve T cell stimulated with antibodies that bind only CD28. 2. A virus enters a cut in the skin of a mouse and infects den- dritic cells, stimulating a variety of PRRs both on and within dendritic cells that induce it to produce IL-12. The mouse subsequently mounts an immune response that successfully clears the infection. Which of the following statements is(are) likely to be true about the immune response that occurred? Correct any that are false. a. The infected dendritic cells up-regulated CD80/CD86 and MHC class II. b. The dendritic cells encountered and activated naïve T cells in the skin of the mouse. c. Naïve T cells activated by these dendritic cells generated signals that released internal Ca2⫹ stores. d. Naïve T cells activated by these dendritic cells were polarized to the TH2 lineage. e. Only effector memory T cells were made in this mouse. 3. Your lab acquires mice that do not have the GATA-3 gene (GATA-3 knockout mice). You discover that this mouse has a difficult time clearing helminth (worm) infections. Why might this be? 4. You isolate naïve T cells from your own blood and want to polarize them to the TH1 lineage in vitro. You can use any of the following reagents to do this. Which would you choose? Anti-TCR antibody IL-4 IFN␥ CTLA-4 Ig anti-CD80 antibody anti-CD28 antibody IL-12 IL-17 e. Signal 3 is provided by negative costimulatory recep- tors. f. Toxic shock syndrome is an example of an autoimmune disease. g. Superantigens mimic TCR-MHC class I interactions. ⫹ ⫹ h. CD4 T cells interact with MHC class I on CD8 T cells. Naïve T cells produce IFN-␥. T-bet and GATA-3 are effector cytokines. Polarizing cytokines are only produced by APCs. Bcl-6 is involved in the delivery of costimulatory signals. m. TH17 and TFH cell subsets are the major sources of B-cell help. n. iTREG cells enhance inflammatory disease. o. Effector cytokines act exclusively on T cells. p. Central memory T cells tend to reside in the site of infection. q. Like naïve T cells, effector memory cells express CCR7. i. j. k. l. 6. Like TH1 and TH2 cells, TH17 and TREG cells cross-regulate each other. Which of these two statements about this crossregulation is (are) true? Correct either, if false. a. TGF-␤ is a polarizing cytokine that stimulates up- regulation of each of the master transcriptional regulators that polarize T cells to the TH17 and TREG lineages. b. IL-6 inhibits polarization to the TREG lineage by inhibiting expression of ROR␥. 7. A new effector T-cell subset, TH9, has been recently identi- fied. It secretes IL-9 and IL-10 and appears to play a role in the protection against intestinal worm infection. What other information about this subset would help you to determine if it should be considered an independent helper T-cell lineage? 12 B-Cell Activation, Differentiation, and Memory Generation T he function of a B cell is to secrete antibodies capable of binding to any organism or molecule that poses a threat to the host. The secreted antibodies have antigen-binding sites identical to those of the receptor molecules on the B-cell surface. Antibodies belong to the class of proteins known as immunoglobulins, and once secreted, they can protect the host against the pathogenic effects of invading viruses, bacteria, or parasites in a variety of ways, as described in Chapters 1 and 13. Our current understanding of B-cell clonal selection, activation, proliferation, and deletion finds its beginnings in a theoretical paper in the Australian Journal of Science, written by Sir Frank MacFarlane Burnet (Figure 12-1) over the course of a single weekend in 1957. In this paper, built on prior work by Neils Jerne, David Talmage, Peter Medawar and others, Burnet laid out the Clonal Selection Hypothesis, which provided the conceptual underpinnings for the entire field of immunology (see Figure 1-7). The Clonal Selection Hypothesis suggested for the first time that the receptor molecule on the lymphocyte surface and the antibody products secreted by that cell had identical antigen-binding specificities. Furthermore, it posited that stimulation of a single B cell would result in the generation of a clone of cells having the identical receptor specificity as the original cell, and that these clones would migrate to, and function within, the secondary lymphoid organs. The daughter cells within each clone would not only be able to secrete large amounts of antibody to neutralize the pathogen, but some progeny cells would also remain viable within the organism and available to neutralize a secondary infection by the same pathogen. In other words, in one brilliant paper Professor Burnet gave generations of immunologists the basis for thinking about B-cell receptor diversity, lymphocyte trafficking, and immune memory—all this at a time when practitioners in the field were still unsure of the differences between T and B cells. Burnet’s paper went on to predict the generation of the vast array of antibody specificities known to exist today. It is difficult for us now, in the second decade of the twentyfirst century, to even begin to appreciate the prescience of The germinal center of the lymph node contains B cells (green) in the dark zone and follicular dendritic cells (red) in the light zone. Naïve B cells (blue) define the follicular mantle zone. [Reprinted from Victora, G.D. et al., 2010, Germinal center dynamics revealed by multiphoton microscopy with a photoactivatable fluorescent reporter, Cell 143:592–605. Copyright 2010 with permission from Elsevier.] ■ T-Dependent B-Cell Responses ■ T-Independent B Cell Responses ■ Negative Regulation of B Cells Burnet’s suggestion that “The theory requires at some stage in early embryonic development a genetic process for which there is no available precedent [italics added]. In some way, we have to picture a ‘randomization’ of the coding responsible for part of the specification of gamma globulin molecules, so that after several cell generations, in early mesenchymal cells, there are specifications in the genomes for virtually every variant that can exist as a gamma globulin molecule.” Just four years after the eludication of the double helical structure of DNA, Burnet was postulating that there are movable genetic segments! The paper ends with a final flourish, in which Burnet predicted the need for clonal deletion in the developing lymphocyte repertoire, in order to eliminate B cells bearing receptors with specificity for self antigens. His formulation of the Clonal Selection Hypothesis, along with the brilliant experimental work he performed with others on the generation of immunological tolerance, resulted in 385 386 PA R T V | Adaptive Immunity: Effector Responses FIGURE 12-1 Sir MacFarlane Burnet, 1899–1985. Professor McFarlane Burnet, the author of the Clonal Selection Hypothesis, shared the 1960 Nobel Prize for Physiology or Medicine with Sir Peter Medawar of Britain for “the discovery of acquired immunological tolerance.” [Sir Macfarlane Burnet 1960–61 by Sir William Dargie. Oil on composition board. Collection: National Portrait Gallery, Canberra. Purchased 1999] his being awarded the Nobel Prize for Physiology or Medicine in 1960, along with Sir Peter Medawar. The essential tenets of the Clonal Selection Hypothesis are summarized below, and a visual representation of the events he predicted are shown in Figure 12-2. • Immature B lymphocytes bear immunoglobulin (Ig) receptors on their cell surfaces. All receptors on a single B cell have identical specificity for antigen. • Upon antigen stimulation, the B cell will mature and migrate to the lymphoid organs. There it will replicate. Its clonal descendants will bear the same receptor as the parental B cell and secrete antibodies with an identical specificity for antigen. • At the close of the immune response, more B cells bearing receptors for the stimulating antigen will remain in the host than were present before the antigenic challenge. These memory B cells will then be capable of mounting an enhanced secondary response. • B cells with receptors for self antigens are deleted during embryonic development. In summary, each B cell bears a single type of Ig receptor generated by the processes described in Chapter 7, and on stimulation will create a clone of cells bearing the same antigen receptor as the original B cell. Many B cells bearing receptors with specificity for self antigens are eliminated from the primary B-cell repertoire as described in Chapter 10. In this chapter, we describe what happens when mature B cells, located in the peripheral lymphoid organs, encounter antigen. The two major types of B-cell responses are elicited by structurally distinct types of antigens. The first type of response that we will describe is generated upon recognition of protein antigens and requires the participation of CD4⫹ helper T cells. Because T cells are involved, this class of B-cell response is therefore known as a T-dependent (TD) response, and it is mediated by B-2 B cells binding to TD antigens. Because B-2 B cells represent the majority of B cells, we will routinely refer to the B-2 B-cell subset simply as “B cells,” and distinguish the other B-cell subclasses by their particular names as B-1, marginal zone, or B-10 B cells. The T-dependent response requires two distinct signals. The first is generated when a multivalent antigen binds and cross-links membrane immunoglobulin receptors (mIg) (Figure 12-3a). The second signal is provided by an activated T cell, which binds to the B cell both through its antigen receptor and via a separate interaction between CD40 on the B cell and CD40L (CD154) on the activated TH cell. The bound T cell then delivers cytokines and other signals to its partner B cell to complete the activation process. The second type of response, which we will describe later in this chapter, is directed toward multivalent or highly polymerized antigens, and does not require T-cell help. This type of response is referred to as a T-independent response, and the antigens that elicit such responses are T-independent (TI) antigens. One class of TI antigens (TI-1 antigens), exemplified by the lipopolysaccharide moiety of Gram-negative bacteria, interacts with the B cell via both mIg and innate immune receptors. TI-1 antigens are mitogenic (induce proliferation) for most B cells at high concentrations, as a result of their ability to bind to pattern recognition receptors (PRRs) on the surface of the B cell. However, at lower concentrations they activate only those B cells that bind antigen with their Ig receptors (Figure 12-3b). The other class of TI antigens, TI-2 antigens, includes highly repetitive antigens, such as bacterial capsular polysaccharides. These antigens are not inherently mitogenic, but their ability to cross-link a large fraction of the Ig receptors on the surface of a B cell allows them to deliver an activation signal in the absence of T-cell help. Many TI-2 antigens are specifically and covalently bound by the complement component C3d, and mice depleted of C3d mount poor responses to TI-2 antigens (Figure 12-3c). Bone marrow Peripheral lymphoid tissue Memory cell 2 IgM IgD 1 Antibody 2 2 2 1 Plasma cells Antigen 2 2 2 2 2 2 2 Gene rearrangement 2 Stem cell 2 3 3 2 2 2 4 Clonal deletion Immature B cells Mature B cells 2 2 Antigen-dependent proliferation and differentiation into plasma and memory cells Maturation into immature committed B cells FIGURE 12-2 Maturation and clonal selection of B lymphocytes. B-cell maturation, which occurs in the absence of antigen, first produces immature B cells bearing IgM receptors. Each B cell bears receptors of one specificity only. Any B cell with receptors specific for antigens expressed in the bone marrow are deleted at the immature B cells stage (indicated by clone 4). Those B cells that do not express self-reactive receptors mature to express both IgM and IgD receptors (a) TD antigen and are released into the periphery, where they recirculate among the blood, lymph, and lymphoid organs. Clonal selection occurs when an antigen binds to a B cell with a receptor specific for that antigen. Clonal expansion of an antigen-activated B cell (number 2 in this example) leads to a clone of effector B cells and memory B cells; all cells in the expanded clone are specific for the original antigen. The effector, plasma cells secrete antibody reactive with the activating antigen. (b) TI-1 antigen Cytokines CD21 TH cell 1 3 1 2 2 B cell (c) TI-2 antigen C3d CD40/CD40L FIGURE 12-3 Different types of antigens signal through different receptor units. (a) T-dependent (TD) antigens bind to the Ig receptor of B cells. Some of the antigen is processed and presented to helper T cells. T cells bind to the MHC-peptide antigen, and deliver further activating signals to the B cell via interaction between CD40L (on the T cells) and CD40 on the B cells. In addition, T cells secrete activating cytokines, such as IL-2 and IL-4, which are recognized by receptors on the B-cell surface. Cytokines deliver differentiation, prolif- B cell TLR 1 B cell eration, and survival signals to the B cells. (b) Type 1 T-independent (TI-1) antigens bind to B cells through both Ig and innate immune receptors. For example, LPS from Gram negative organisms binds to B cells via both mIg and TLR4, resulting in signaling from both receptors. (c) Type 2 T-independent (TI-2) antigens are frequently bound by C3d complement components and cross-link both mIg and CD21 receptors on B cells. Cross-linking of between 12 and 16 Ig receptors has been shown to be sufficient to deliver an activating signal. 388 PA R T V | Adaptive Immunity: Effector Responses Most T-independent responses are mediated by B-1 and marginal zone B-cell types. T-Dependent B-Cell Responses Following the completion of their maturation program (see Chapter 10), B cells migrate to the lymphoid follicles (Figure 2-8). They are directed there by chemokine interactions between CXCL13, which is secreted by follicular dendritic cells (FDCs) and its receptor, CXCR5, which is expressed on B cells. It is important to recognize, however, that once mature B cells have reached the follicles, they do not just remain there; rather, they recirculate through the blood and lymphatic systems and back to the lymphoid follicles many times over the course of their existence. B-cell survival in the follicle is dependent on access to the TNF-family member cytokine B-cell activation factor (BAFF), otherwise known as B lymphocyte survival factor (BLyS), which is secreted by the FDCs, as well as by many types of innate immune cells, such as neutrophils, macrophages, monocytes, and dendritic cells. Mature B cells unable to secure a sufficient supply of BAFF die by apoptosis. Recirculating, mature B cells are estimated to have a half-life of approximately 4.5 months. At the initiation of a T-dependent B-cell response, the B cell binds antigen via its Ig receptors. Some of that bound antigen is internalized into specialized vesicles within the B cells, where it is processed and re-expressed in the form of peptides presented in the antigen-binding groove of class II MHC molecules (see Figure 8-19). T cells that have been previously activated by an encounter with antigen-bearing dendritic cells now bind to the MHC-presented peptide on the surface of the B cell. This induces a vectorial redistribution of the T cell’s secretory apparatus, such that it releases its activating cytokines directly into the T-cell/B-cell interface (as described in Chapter 4). Other interactions also occur between accessory molecules on the T- and B-cell surfaces that enhance their binding and provide further signals to the pair of cells. For example, CD40L on the T cell binds to CD40 on the B cell (see Figure 12-3a) and the CD28 coreceptor on the T cell engages with CD80 and CD86 on the surface of the B cell. The B cell then integrates the activating signals received through its antigen receptor with those from its various cytokine receptors and coreceptors and moves into specialized regions of the lymph node or spleen to begin the process of differentiation into an antibody-secreting cell. Recall that lymphocytes first enter the lymph node in the region of the T-cell zone, just inside the follicles (Figure 2-8). Some of the antigenactivated B cells then move into regions at the borders of the T-cell and B-cell areas, where they differentiate into clusters of activated B cells known as primary foci (see Overview Figure 12-4). There, they complete their differentiation into plasma cells. At around 4 days post stimulation, when this differentiation is complete, they migrate into the medullary cord regions of the lymph node, or to parts of the red pulp of the spleen close to the T-cell zones (Figure 2-10), where they secrete large quantities of antibodies. Some of these plasma cells die after the initial primary response is completed, whereas others take up long-term residence in the bone marrow, gut, or other locations as long-lived plasma cells. These primary focus plasma cells deliver relatively large quantities of IgM antibodies in the early phases of the B-cell response. Some antigen-stimulated B cells do not enter the primary foci but rather migrate into the follicles of the lymph nodes and spleen where they undergo further differentiation. As the follicles swell with antigen-specific lymphocytes (primarily B cells), they change their appearance and become known as germinal centers (GCs). In these germinal centers, B-cell variable region genes undergo mutational processes that result in the secretion of antibodies with altered sequences in their antigen-combining sites. This process of somatic hypermutation (SHM) is then followed by antigen selection, culminating in the production of B cells bearing receptors and secreting antibodies whose affinity for antigen increases as the immune response progresses. Antibodies with mutations in their variable regions begin to appear in the circulation 6 to 10 days following the onset of the immune response. In addition, ongoing signals from helper T cells direct their cognate B cells to produce antibodies of isotypes (classes) other than IgM, in a process known as class switch recombination (CSR). Both SHM and CSR are dependent on the activity of a germinal center enzyme: activationinduced cytidine deaminase (AID). The end result of this extraordinary set of events is the production of high-affinity antibody molecules that eliminate the pathogenic threat by one or more of the means described in Chapter 13. At the close of the primary immune response, memory B cells remain that are the daughters (and more distant progeny) of those cells that were stimulated during the primary response. Many of these progeny B cells now carry mutated and selected B-cell receptors (BCRs). On secondary exposure to the same antigen, these memory B cells will be stimulated more quickly and will deliver higher-affinity and heavy-chain class-switched antibodies, resulting in faster elimination of microbial pathogens. This improved memory response underlies the scientific rationale for vaccination. As the immune response to T-dependent antigens winds down, the host organism is left with two sets of long-lived cells that ensure the provision of long-term memory responses to the antigen. Plasma cells in the bone marrow and elsewhere create a reservoir of antibody-producing cells that can last for the lifetime of the host, and memory B cells circulating through the lymphoid organs remain poised for subsequent stimulation by the same antigen. Excess residual lymphocytes and plasma cells from the primary response are eliminated by apoptosis. T-Dependent Antigens Require T-Cell Help to Generate an Antibody Response The initial experiments proving that B cells required “help” from T cells in order to complete their differentiation were performed by Miller, Mitchell, Mitchison, and B-Cell Activation, Differentiation, and Memory Generation | CHAPTER 12 389 12-4 OVERVIEW FIGURE B-Cell Activation T cell zone Germinal center Differentiation Selection Low affinity Apoptosis Proliferation, SHM, and CSR Conventional naïve B cell TD antigen activation Plasma cell FDC Class-switched memory B cell Improved affinity TD antigen activation Abortive autoreactive GC B cell TH cell Apoptosis IgM+ IgD+ memory B cell Primary focus Dark zone Light zone Medullary cords B cells engaged in T-dependent activation first encounter antigenspecific T cells outside of the B-cell follicles. Some activated B cells differentiate into antibody-producing plasma cells in primary foci that lie outside the follicles, and then migrate to the medullary cords of the lymph node, or the bone marrow, where they continue to secrete antigen-specific antibodies. Other antigen-activated B cells enter the follicle where they divide and differentiate. As the follicle fills with proliferating B cells, it develops into a germinal center, characterized by a light zone, in which follicular dendritic others in the 1960s, using the technique of adoptive transfer. In an adoptive transfer experiment, a mouse is lethally irradiated in order to eliminate its immune system; the investigator then attempts to reconstitute the ability to generate an immune response by adding back purified cells of different types from genetically identical mouse donors. Using this technique, investigators showed that, in order for a mouse to produce antibodies against a protein antigen, it must receive cells derived from both the bone marrow and from the thymus of a healthy donor animal. Neither thymusderived nor bone marrow-derived cells were capable of reconstituting the responding animal’s immune system on their own (Figure 12-5). We now know, of course, that the thymusderived cells active in this response were mature helper T cells, cells reside, and a dark zone in which the cell density is particularly high. Within the germinal center, the immunoglobulin genes undergo class switching, in which ␮ constant regions are replaced by constant regions of other isotypes. The variable region is subject to somatic hypermutation, and the mutated variable regions are subject to antigen-mediated selection within the germinal centers. Low-affinity and autoimmune receptor-bearing B cells die, and those B cells with enhanced receptors leave the germinal centers for the periphery. (See the text for details.) whereas the bone marrow-derived, antibody-producing cells were mature B cells, recirculating through the bone marrow. In this way it was demonstrated that the antibody response to protein antigens required both B cells and T cells. Having introduced the major players and briefly described the geography of the landscape in which the cells are operating, we will now sequentially step through a T-dependent B-cell response. Antigen Recognition by Mature B Cells Provides a Survival Signal The first step in antibody production is antigen recognition by the Ig receptors on the surface of a naïve B cell (a B cell 390 PA R T V | Adaptive Immunity: Effector Responses 1. Irradiated recipient mice 2. Reconstituted recipient mice with: Nothing 3. Measured ability of recipient mice to mount an antibody response to a T-dependent antigen (sheep red blood cells) Bone marrow cells alone –– –– Thymus cells alone –– Bone marrow and thymus cells ++ FIGURE 12-5 Adoptive transfer experiments demonstrated the need for two cell populations during the generation of antibodies to T-dependent antigens. Early adoptive transfer experiments reconstituted syngeneic irradiated mice, in which the immune system had been ablated, with bone-marrow cells, thymus-derived cells, or a mixture of bone-marrow and thymus-derived cells. These mice were then challenged with a T-dependent antigen. Only recipient mice reconstituted with both bone-marrow and thymus-derived cells were able to mount an antibody response. that has not yet encountered antigen). These receptors, along with their coreceptors and signaling properties, were described in Chapter 3. Naïve B cells circulate in the blood and lymph and pass through the secondary lymphoid organs, most notably to the spleen and lymph nodes (Chapter 2). If a B cell is activated by interaction with an antigen, the cell will proliferate and differentiate as described below. If a B cell does not bind antigen, it will die by apoptosis a few months after its emergence from the bone marrow. B Cells Encounter Antigen in the Lymph Nodes and Spleen When antigen is introduced into the body, it becomes concentrated in various peripheral lymphoid organs. Bloodborne antigen is filtered by the spleen, whereas antigen from tissue spaces drained by the lymphatic system is filtered by regional lymph nodes (or lymphoid nodules in the gut). Here, we will focus on antigen presentation to B cells in the lymph nodes. Antigen enters the lymph nodes either alone or associated with antigen-transporting cells. Recall that B cells are capable of recognizing antigenic determinants on native, unprocessed antigens, whereas T cells must recognize antigen as a peptide presented in the context of MHC antigens. In addition, as described in Chapter 6, antigen is often covalently modified with complement fragments, and complement receptors on B cells play an important role in binding complement-coupled antigen with sufficient affinity to trigger B-cell activation. The mechanism of B-cell antigen acquisition varies according to the size of the antigen. Small, soluble antigens can be directly acquired from the lymphatic circulation by follicular B cells, without the intervention of any other cells. These antigens enter the lymph node via the afferent lymph and pass into the subcapsular sinus region (Figure 12-6a,b). Some small antigens may diffuse between the subcapsular sinus (SCS) macrophages that line the sinus to reach the B cells in the follicles. Others leave the sinus through a reticular (netlike) network of conduits (cylindrical vessels). These conduits are leaky, and follicular B cells can access antigen through gaps in the layer of follicular reticular cells that form the walls of the conduits (Figure 12-6b). Larger, more complex antigens take a slightly different route. The SCS macrophages, shown just beneath the subcapsular sinus in Figure 12-6a, are a distinct subpopulation of lymph node macrophages with limited phagocytic ability, and express high levels of cell-surface molecules able to bind and retain unprocessed antigen. For example, bacteria, viruses, particulates, and other complex antigens that have been covalently linked to complement components are held by complement receptors on the surfaces of these macrophages. Antigen-specific B cells within the follicles can acquire the antigens directly from the macrophages and become activated. In addition, since SCS macrophages, B cells, and FDCs all bear high levels of complement receptors, it is quite possible that antigen may be passed between the three subtypes of cells until it is finally recognized by a B cell with the matching Ig receptor. Other classes of cells within the lymph node can also present unprocessed antigens to B cells. For example, a population of dendritic cells that is located close to the high endothelial venules in lymph nodes has also been shown to be capable of presenting unprocessed antigen to B cells. Any discussion of antigen presentation to B cells would be incomplete without consideration of the role of follicular dendritic cells (FDCs). Early electron microscope investigations of lymph-node-derived cells revealed the dendrites of FDCs to be studded with antigen-antibody complexes called iccosomes (Figure 12-6c). Further analysis showed that these complexes are retained on the surface of the FDC through interaction either with complement or with Fc receptors. Because of the high surface density of antigen on FDCs, many scientists at first posited that iccosomes’ primary role may be in antigen presentation to naïve B cells. However, current evidence suggests that their main function | B-Cell Activation, Differentiation, and Memory Generation (a) Subcapsular sinus CHAPTER 12 391 (b) Subcapsular sinus pore Afferent lymph Chemokine Small antigen (<70 kDa) Large antigen (>70 kDa) SSM Follicular conduit Follicular B cell FRC Antigen specific B cell Follicular dendritic cell IgM Lymph IgD Collagen Antigen non-specific B cell Follicle FIGURE 12-6 Antigen presentation to follicular B cells in (c) the lymph node. (a) Lymphatic fluid containing antigens (red) and cytokines and chemokines (blue) reaches the lymph node through the afferent lymph vessel and enters the subcapsular sinus (SCS) region. The SCS region is lined with a porous border of SCS macrophages (SSMs) that prevent the free diffusion of the lymph fluid into the lymph node. Larger antigens and signaling molecules are bound by surface receptors on the SCS macrophages and then presented directly to B cells. Smaller antigens, and chemokines less than approximately 70 kDa in molecular weight, access B cells in the follicles either by direct diffusion through pores in the SCS, or by passage through conduits emanating from the sinus. (b) These conduits are formed by follicular reticular cells (FRCs) wrapped around collagen fibers, and B cells are able to access their contents through pores in the sides of the conduits. (See the text for further details.) (c) A follicular dendritic cell with its dendrites studded with iccosomes. Follicular dendritic cells bind antigen-antibody complexes via complement or Fc receptors on their cell membranes. These complexes are visible in the electron microscope, as iccosomes. [Parts a and b adapted from Harwood, N.E., and Batista, F.D., 2009, The antigen expressway: follicular conduits carry antigen to B cells, Immunity 30:177–189. Part (c) Courtesy Andras K. Szakal PhD.] is to provide a reservoir of antigen for B cells to bind as they undergo mutation, selection, and differentiation along the path to a memory phenotype, rather than playing the primary role in the initial antigen presentation process. In addition, FDCs secrete survival factors that ensure the survival of B cells within the lymph node. B-Cell Recognition of Cell-Bound Antigen Results in Membrane Spreading Prior to contact with antigen, the majority of BCRs are expressed in monomeric, bivalent form on the cell surface. Interaction of these monomeric receptors on the surface of the B cell with multivalent, cell-bound antigens then induces a rather spectacular cellular response. First, a cluster of BCRs and their cognate antigens forms at the initial site of contact. Next, the B cell rapidly spreads over the target membrane, before contracting back. This spreading reaction can be clearly seen in Figure 12-7, in which the cell-bound antigen was modeled using a protein incorporated into a planar lipid bilayer. Membrane spreading occurs 2 to 4 minutes after antigen contact, during which time micro-clusters of antigen-receptor interactions can be seen by time-lapse fluorescence microscopy at the membrane-lipid interface. After reaching a maximum surface area of approximately 25 ␮m2, the area of contact between the cell and the artificial lipid membrane begins to contract, and the antigen-receptor complex is gathered into a central, defined PA R T V 392 | Adaptive Immunity: Effector Responses By the end of the contraction phase, the BCR, still in contact with its cognate antigen on the presenting cell, is clustered on the surface of the B cell. What Causes the Clustering of the B-Cell Receptors Upon Antigen Binding? FIGURE 12-7 Antigen recognition by the BCR triggers membrane spreading. Transgenic B cells expressing a BCR specific for hen egg lysozyme (HEL) were settled onto planar lipid bilayers containing HEL and were incubated for the time period shown, followed by fixation and visualization with scanning electron microscopy. At 2 to 4 minutes, the B-cell membrane can clearly be seen spreading over the surface of the planar lipid bilayer (arrows). By 5 minutes, the membrane begins to contract again, after which the BCR molecules are found clustered on the cell surface. (See the text for details.) [Fleire, S.J., Goldman, J.P., Carrasco Y.R., Weber, M., Bray, D. and Batista, F.D. B cell ligand discrimination through a spreading and contraction response. Science 5 May 2006, Vol. 312, no. 5774, pp. 738–741. © 2006 The American Association for the Advancement of Science] cluster with an area of approximately 16 ␮m2. The contraction phase takes an additional 5 to 7 minutes. The effectiveness of the ultimate antibody response has been shown to correlate with the extent of this spreading reaction. Experiments using monovalent ligands incorporated into target lipid membranes demonstrated that the antigen itself need not be multivalent in order for clustering to begin. Rather, it appears that antigen binding to the BCR at sufficiently high affinity causes a structural alteration in the BCR, rendering it susceptible to clustering. Initially, antigen ligation of the Ig receptor is associated with a decrease in the rate of diffusion of IgM receptors in the plane of the membrane. In addition, antigen binding appears to induce a conformational change in the C␮4 constant region domains of occupied IgM membrane receptors. These two changes result in an increase in the tendency of the bound IgM receptors to bind to the corresponding domains of other antigen-bound IgM receptor molecules (Figure 12-8). The deceleration in the receptor’s movement in the membrane coupled with this receptor oligomerization occurs whether or not the Ig␣,Ig␤ (CD79␣,␤) signaling components of the receptor are present, indicating that the receptor clustering does not require signaling events. It is now thought that this BCR clustering may be the initiating event in antigen signaling through the BCR. Upon oligomerization, the BCR complex moves transiently into parts of the membrane characterized by highly ordered, detergent-insoluble, sphingolipid- and cholesterolrich regions, colloquially designated as lipid rafts. Recall that association of the BCR with the lipid rafts then brings the ITAMs of the Ig␣ and Ig␤ components into close apposition with the raft-tethered src-family member tyrosine Planar lipid membrane mIgM Monovalent antigen Cµ4 domain Igα/β Lyn Lyn B cell membrane Antigen-binding B cell FIGURE 12-8 Antigen binding induces a conformational change in the C␮4 domain and subsequent oligomerization of antigen-bound IgM molecules in the plane of the membrane. When monovalent antigens floating in an artificial bilayer encounter IgM receptors, they bind and induce a conformational change in the C␮4 domain of the IgM heavy chain. This conformational change facilitates oligomerization of the IgM receptors in the plane of the membrane. (See the text for details). [Adapted from P. Tolar et al., 2009, The molecular assembly and organization of signaling active B-cell receptor oligomers, Immunological Reviews 232:34–41, Figure 2.] B-Cell Activation, Differentiation, and Memory Generation kinase, Lyn. This sets off the B-cell signaling cascade shown in Figure 3-28. Antigen Receptor Clustering Induces Internalization and Antigen Presentation by the B Cell Antigen binding by the B cell activates the signaling cascade described in Figure 3-28, and this cascade induces several changes within the cell. One of these changes is the formation around the BCR of clathrin-coated pits that facilitate the internalization of the receptor-antigen complexes. Indeed, we now know that most of the antigen-occupied BCR molecules are internalized, leaving just a few copies of the receptor on the cell surface to serve as signaling scaffolds. Experiments have shown that those antigen-bound BCR-Ig␣,Ig␤ receptor complexes that remain at the cell surface and provide the B-cell signaling function are more highly phosphorylated than are those receptors that are internalized. Once internalized, the antigen is processed as described in detail in Chapter 8. This results in an increased expression of peptide-loaded class II MHC molecules on the B-cell surface, as well as an up-regulation of the expression of the costimulatory molecules CD80 and CD86. Recall that CD80 and CD86 both bind to the T-cell costimulatory molecule CD28. These changes in the expression of CD80 and CD86, which occur within 1 to 2 hours after antigen recognition, therefore prepare the B cells for their subsequent interactions with T cells. A B cell that has taken up its specific antigen via receptormediated endocytosis is up to 10,000-fold more efficient in presenting antigen to cognate T cells than is a nonantigenspecific B cell that can only acquire the same antigen by nonspecific pinocytotic mechanisms. Effectively, this means that the only B cells that process and present antigen to T cells in a physiologically relevant context are those B cells with the capacity to make antibody to that antigen. Now that the B cell is ready to present antigen to T cells, we must ask how a B cell, with a receptor that is normally expressed at extremely low frequency within the receptor repertoire, can possibly find and bind to a T cell specific for the same antigen, which is also present at low frequency among all the available T cells? Activated B Cells Migrate to Find Antigen-Specific T Cells We owe a great deal of our understanding of the movement of cells and antigens through the lymph nodes to recent advances in imaging cells within their biological context. Specifically, in some recent experiments described in Chapter 14, lymph nodes have been brought outside the body of living, anesthetized animals, and lymphocyte circulation through these nodes has been studied using fluorescently tagged T, B, and antigen-presenting cells. In these experiments, the blood and lymphatic circulation to the lymph node is preserved, and the lymph node is gently lifted onto a warmed, humidified | CHAPTER 12 393 microscope stage for observation. Sometimes the fluorescent tags are built into the genome of the animal by placing fluorescent proteins, such as green fluorescent protein (GFP), under the control of specific promoters that are active only in specific lymphocyte subsets. In other experiments, the animal is injected with fluorescent antibodies just prior to being anesthetized. This technique is known as intravital fluorescence microscopy (see Chapters 14 and 20) and has contributed hugely to our understanding of the process of lymphocyte activation in the lymph nodes. Investigators have used immunohistochemical, immunofluorescence, and the aforementioned intravital imaging techniques to visualize particular cell populations during the course of an ongoing immune response. This has, in some cases, required the generation of transgenic mice in which all, or most, B and T cells have defined antigen specificity. All of these techniques are described in Chapter 20. What follows is a distillation of the information gained from a large number of such experiments using various antigen and transgenic model systems. As we have learned, B cells pick up their antigen in the follicular regions of the lymph node or spleen. By approximately 2 hours post antigen contact, the B cell has internalized and processed its antigen and expressed antigenic peptides on its surface in the context of class II MHC antigens. In response to antigen recognition by the BCR, the B cell begins to express the chemokine receptor CCR7. CCR7 binds to the chemokines CCL19 and CCL21, which are secreted by stromal cells in the T-cell zones of secondary lymphoid organs. The B cell still also expresses CXCR5, which binds CXCL13, expressed by FDCs in the B-cell follicles. Because they express receptors for both T- and B-cellzone-derived chemokines, by approximately 6 hours post stimulation, antigen-engaged B cells move to the boundary of the B-cell and T-cell zones. Figure 12-9a shows these various lymph node zones in diagrammatic form. Activated B cells also up-regulate the expression of receptors for cytokines released by activated T cells, allowing B cells to receive the signals to proliferate, to differentiate, and also to initiate an anti-apoptotic program. The receptor for IL-4, an important B-cell-specific cytokine, can be detected as early as 6 hours post antigen contact and reaches maximal levels at 72 hours. Antigen-stimulated B cells move about within the T-cell zone of the lymph node until they make contact with an antigen-specific T cell, an event that most probably occurs over the 24 to 48 hours post antigen contact. By this time, many of the important cell-surface changes characteristic of B lymphocyte activation can be observed, such as increased levels of class II MHC antigens and of the two costimulatory molecules, CD80 and CD86. Once contact with an activated T cell is made, these antigen-stimulated B cells engage with their conjugate T-cell partners over extended periods of time, ranging from a few minutes to several hours. During this period, the activated helper T cell also expresses CD40L (CD154), a cell-membrane-bound member of the TNF receptor family, which interacts with CD40 (a TNF family member) 394 PA R T V | Adaptive Immunity: Effector Responses (a) (b) Naïve T-cell zone T-B border Germinal center Inter-follicular zone Sub-SCS Follicle Day 1 FIGURE 12-9 Movement of antigen-specific T and B cells within the lymph node after antigen encounter. (a) Section of the lymph node in diagrammatic form, showing the various regions of the lymph node. (b) The locations of antigen-specific B cells (green) and antigen-specific T cells (red) within the lymph node were visualized at specified times after antigen stimulation. The antigen used in the experiment was nitro-phenacetyl (NP), conjugated to ovalbumin (OVA). Fluorescently labeled NP-specific B cells and OVA-specific T cells were transferred into a recipient animal, which was immunized with NP-OVA in the footpads 2 days after cell transfer. Draining popliteal lymph nodes were excised at the stated times. Anti-B220 (blue) was used to identify the B-cell follicles. The times refer to the days (D) post immunization. D1 and D2: Antigen-specific B-T cell pairs could be found both at the border of the T- and B-cell zones and in the interfollicular regions. D3: T cells have begun to enter the follicle, and many B cells can be seen just underneath the subcapsular sinus of the lymph node. D4: B cells have taken up residence in the follicle, and the formation of the germinal center can be seen. [Fleire, S.J., Goldman, J.P., Day 2 Day 3 Carrasco Y.R., Weber, M., Bray, D. and Batista, F.D. B cell ligand discrimination through a spreading and contraction response. Science 5 May 2006, Vol. 312, no. 5774, pp. 738–741. ©2006 The American Association for the Advancement of Science] (see Chapter 4) on the activated B cell. The CD40-CD40L interactions, coupled with BCR-mediated antigen recognition and interleukin signaling, elicit all of the downstream activation pathways shown in Figure 3-28, and together drive the B cell into the proliferative phase of the cell cycle. After this period of intense communication between the activated T and B cells, activated B cells then down-regulate CCR7 while maintaining high levels of expression of CXCR5. This allows them to leave the T-cell areas and to migrate into the inter- and outer-follicular regions. The activated B cells appear to spend 1 or 2 days in the regions to the outside and in between the follicles before finally entering the follicle around 4 days post immunization (Figure 12-9b). In the experiment depicted in Figure 12-9b, a transgenic mouse has been immunized with the T-dependent antigen nitrophenacetyl-ovalbumin (NP-OVA). Antigen-specific B cells express GFP, and are therefore labeled green; antigenspecific T cells are labeled red; and the B-cell follicles are stained blue. In the naïve animal, T cells can be seen clearly localized in the T-cell zones, and occasional green B cells can be seen scattered throughout the follicles. With time post immunization, B and T cells can be seen migrating first into the interfollicular regions between the follicles (days 1 and 2 post immunization). By day 3, B and T cells can be found Day 4 B-Cell Activation, Differentiation, and Memory Generation clustered in the outer- as well as the interfollicular regions, and by 4 days post immunization most, but not all, of the B cells have entered the follicles and the beginning of germinal center development can be seen. Interestingly, B cells derived from mice deficient in CXCR5 or CXCL13 behave in a fashion that is not as different from those derived from normal mice as might have been expected, and this observation led to the discovery of a second ligandreceptor pair that induced the same migrating behavior in B cells. EBI2-ligand is recognized by a G-protein-coupled receptor, EBV-induced ligand 2 receptor (EBI2), that is present on all mature B cells. The level of expression of EBI2 increases markedly and quickly following B-cell activation, and interaction of EBI2 with its ligand results in the movement of B cells into the outer- and interfollicular regions in a manner analogous to that instructed by CXCR5 interaction with CXCL13. Biochemical experiments suggest that the ligand for EBI2 is lipid in nature, and investigators are still working to determine the nature of the cell types that express the ligand, the dynamics of its expression, and the way in which the two sets of chemoattractant signals interact to ensure the correct movement of the antigen-stimulated B cell through the lymph node. The outer- and interfollicular regions of the lymph node contain high numbers of SCS macrophages, dendritic cells, and a cell population referred to as marginal reticular cells (MRCs), and the roles of each of these cell populations in the migratory behavior of antigen-activated B cells is another area of active investigation. Activated B Cells Move Either into the Extrafollicular Space or into the Follicles to Form Germinal Centers Over the first few days after their interaction with T cells, daughter cells of the proliferating B cell elect one of two fates. Some of the stimulated daughter cells in the extrafollicular spaces migrate into the borders of the T-cells zone, form a primary focus, and differentiate into plasmablasts. Plasmablasts are B cells that can still divide and present antigen to T cells, but they have already begun to secrete antibody. Plasmablasts and plasma cells in the primary focus secrete measurable levels of IgM by approximately 4 days after antigen contact and are responsible for the earliest manifestations of the antibody response. Other daughter cells generated by stimulation of the original B cells enter the B-cell follicles, as shown in Figure 12-9b in the bottom panel. There, they divide rapidly and undergo further differentiation. This movement of B cells into the follicles is facilitated by the expression of the transcription factor Bcl-6, which represses the expression of EBI-2 and enables the B cells to leave the outer- and interfollicular regions. As the B cells differentiate under the influence of follicular helper T cells (TFH) (see Chapter 11), the follicle becomes larger and more dense, developing into a germinal center. (The germinal center reaction will be discussed below.) The precise nature of the mechanisms responsible for determining which cells enter the primary focus and which | CHAPTER 12 395 elect to enter the follicle and establish a germinal center reaction are unclear at this time and may be stochastic. Below, we first describe the processes that lead to primary focus formation and then go on to discuss the extraordinary fate that awaits those B cells that enter the follicles, and eventually develop into germinal center B cells. Plasma Cells Form within the Primary Focus Plasma cells are essentially Ig-producing machines. Their surface Ig levels are close to zero, and they are no longer capable of being further stimulated by antigen, or of presenting antigen to responsive T cells. An activated B cell in the extrafollicular regions of the lymph node that is initiating a program of differentiation toward the plasma cell endpoint begins to divide rapidly and to secrete low levels of Ig. As the cell divides, it decreases levels of membrane Ig, MHC proteins, and CD80/86, and increases its rate of Ig secretion to become a plasmablast. Eventually, the cell reaches a stage of terminal differentiation in which it is no longer capable of cell division and has achieved a maximal rate of Ig secretion: the plasma cell. Plasma cells are found within the first 5 to 6 days of an immune response in the medullary cord region of the lymph node. Most primary focus plasma cells have short half-lives, dying by apoptosis within 5 to 10 days of their generation, and for many years it was thought that all plasma cells from the primary foci endured this fate. However, recent experiments have suggested that some of these plasma cells may migrate to the bone marrow or to other locations within the body, where they provide long-lasting Ig memory. There is currently a great deal of interest in the transcription factors that control whether antigen-stimulated B cells differentiate along the plasma cell or, alternatively, the germinal center route. Scientists now understand that these transcription factors are linked in a mutually regulatory network (Figure 12-10). Pax-5 IRF-4 Bcl-6 BLIMP-1 Germinal center fate Plasma cell fate SHM CSR FIGURE 12-10 A regulatory network of transcription factors controls the germinal center B cell/plasma cell decision point. The transcription factors that control germinal center B cell versus plasma cell states of differentiation are related to one another through a mutually regulatory network. (See the text for details). 396 PA R T V | Adaptive Immunity: Effector Responses Transcription factors that favor the generation of proliferating germinal center B cells, Pax-5 and Bcl-6, actively repress the expression of the plasma cell transcription factor BLIMP-1. Indeed, bcl-6 gene knockout animals are incapable of forming germinal centers, a factor that demonstrates the centrality of Bcl-6 in determination of the germinal center phenotype. Immune responses from these knockout mice also fail to undergo affinity maturation and class switch recombination, processes that depends on the normal functioning of germinal centers. In wild-type animals, as illustrated in Figure 12-10, induction of BLIMP-1 and IRF-4 results in the inhibition of Bcl-6 and Pax-5 expression. This antigen-induced expression of two alternative sets of transcription factors that control the decision between two potential lymphocyte cell fates is reminiscent of the induction of T-Bet and GATA-3 and their respective effects on TH1 and TH2 induction (Figure 11-10). In both cases, induction of the transcription factor(s) that initiate one differentiation pathway results in the direct inhibition of the synthesis and activity of the transcription factors supporting the alternative pathway. The transcription factor B-lymphocyte-induced maturation protein 1 (BLIMP-1) had long been considered to be the master regulator of plasma cell differentiation. BLIMP-1 knockout animals lack all capacity to generate plasma cells, while enforced expression of BLIMP-1 is sufficient to promote plasma cell differentiation. In wild-type animals, BLIMP-1 expression peaks at the plasma cell stage of B-cell differentiation, and the highest levels of BLIMP-1 expression are seen in long-lived plasma cells in the bone marrow and primary focus plasma cells in the secondary lymphoid organs. Expression of BLIMP-1 also decreases the levels of MHC expression on the surface of the cell, consistent with the inability of plasma cells to engage in presentation of antigen to T cells. BLIMP-1 also promotes the alternative splicing of Ig mRNA, which enables the generation of mature transcripts encoding the secreted form of antibodies (see Figure 7-15). However, when scientists asked whether BLIMP-1 was the first transcription factor operative in the cellular process that induces B cells to differentiate into plasma cells, the answer was not as straightforward as had been expected. Specifically, recent experiments have demonstrated that some genes associated with plasma-cell formation are still turned on in blimp-1 knockout mice, suggesting that some factor other than BLIMP-1 may initiate the process of plasma cell differentiation, even if BLIMP-1 is required for its completion. What other factor(s) might be responsible for initiating the plasma-cell fate pathway? Current evidence points to the Interferon Regulatory Factor 4 (IRF-4) as the most probable candidate (see Figure 12-10). If IRF-4 is controlling the expression of blimp-1, we would expect it to be expressed before BLIMP-1 in the developing B cell, and to control the expression of the blimp-1 gene. Such is indeed the case. The irf-4 gene is expressed before blimp-1, and the IRF-4 protein binds to elements upstream of the blimp-1 gene, up-regulating its transcription. Furthermore IRF-4 down-regulates the expression of the genes encoding both Pax-5 and Bcl-6. Finally, the concentration of IRF-4 varies in B cells at different stages of differentiation in a manner consistent with it being a determining factor in plasma cell differentiation. Thus, it appears that IRF-4 may initiate differentiation to the plasma cell stage. But is it required, or merely a useful adjunct to the process of plasma-cell differentiation? For the answer to this question, we must turn to knockout experiments. Experiments with IRF-4-deficient mice have shown that they completely lack Ig-secreting plasma cells; conversely, and as for BLIMP-1, overexpression of IRF-4 promotes plasma cell differentiation. IRF-4 has been shown to bind to Ig enhancer regions and is important in the generation of high levels of Ig secretion. Thus, it would appear that IRF-4 may be a determining factor, or possibly the determining factor, in driving a B cell toward the plasma cell, rather than the germinal center phenotype. The generation of the plasma cells in the primary focus is a critically important step in the antibody response, and the rate of secretion of Ig by these cells is extraordinarily high. At the peak of an antibody response, a single antibody-secreting cell can release as much as 0.3 ng of Ig per hour, with fully 30% of its cellular protein synthesis devoted to the generation of secreted antibody molecules. Therefore, these cells provide high levels of antibody that can bind antigen, neutralizing it and/or opsonizing it for phagocytosis by macrophages. However, the antigen-binding affinities of the antibodies secreted by the primary focus have not yet been optimized by the related processes of somatic hypermutation and antigen selection, and their affinity for the antigen is often relatively low. In the case of a particularly virulent infection, the antibodies secreted by the cells of the primary focus may therefore serve primarily to contain microbial numbers until the high-affinity antibodies generated by the germinal center reaction are released. It is within the germinal center that the next extraordinary events in B-cell differentiation will occur. Other Activated B Cells Move into the Follicles and Initiate a Germinal Center Response As described above, 4 or 5 days after T-dependent B-cell activation occurs at the border between the T-cell zone and the follicles, some antigen-specific B cells enter the follicles. There, they undergo rapid proliferation, resulting in the formation of large clusters of antigen-specific B cells. The follicle, which contains both TFH and follicular dendritic cells (FDCs), becomes larger as the entering antigen-specific cells proliferate, and the teeming follicle in which an immune response is actively ongoing is referred to as a germinal center (GC) (Figure 12-11). Interaction between CD40 on the B cell and its ligand, CD40L, on the T cell is necessary for germinal center formation, and no germinal centers are generated in mice in which the genes for either CD40 or CD40L have been deleted. In the germinal centers, B cells undergo a period of intense proliferation, and then their Ig genes are subjected to some of B-Cell Activation, Differentiation, and Memory Generation (a) | CHAPTER 12 397 (b) Follicular mantle zone Light zone Dark zone Dark zone Light zone FIGURE 12-11 The germinal center. (a) A mouse lymph node germinal center 7 days after secondary immunization, in which 10% of the germinal center B cells express green fluorescent proteins (and are therefore labeled green). These proteins are seen most intensely in the dark zone of the germinal center, but are also present less frequently in the light zone. Follicular dendritic cells, labeled red with an antibody to the FDC marker CD35, mark the light zone of the germinal center. The blue staining marks IgDthe most extraordinary processes in biology. First, the Ig variable (but not the constant) region genes undergo extremely high rates of mutagenesis, on the order of one mutation per thousand base pairs per generation. (Contrast this with the background cosmic ray-induced mutation rate of one mutation per hundred million base pairs per generation.) The mutated genes are then expressed, and the encoded Ig receptors are tested to see if their affinities for antigen have been altered for the better. Those B cells bearing higher-affinity receptors than those on the parental cells are selected for further rounds of mutation and selection, while those with no, or decreased, affinity for antigen are allowed to die by apoptosis. Second, also within the germinal center, interleukin signals from follicular T cells drive the process of Ig CSR: the replacement of the ␮ constant region gene segments with segments encoding other classes of constant regions. We are just beginning to understand the molecular biology of these processes, which take place in a rapidly modulating anatomical environment. Given the active mutational and selection events that occur within the germinal center, it should come as no surprise that it has been referred to by some as a “Darwinian Microcosm.”1 Below, we will describe the biology of the germinal center and then look in turn at the two genetic processes that are unique to the immune system: somatic hypermutation and the Ig class switch. 1 Garnett Kelsoe. (1998). V(D)J hypermutation and DNA mismatch repair: Vexed by fixation. Proceedings of the National Academy of Sciences of the United States of America 95:6576. bearing B cells not specific for the immunizing antigen. (b) Histochemical staining of a germinal center, illustrating the remarkably high centroblast cell density within the dark zone. [(a) Reprinted from Victora, G.D. et al., 2010, Germinal center dynamics revealed by multiphoton microscopy with a photoactivatable fluorescent reporter, Cell 143:592–605. Copyright 2010 with permission from Elsevier. (b) Courtesy Dr. Roger C. Wagner, Professor Emeritus of Biological Sciences, University of Delaware] Germinal Center Formation When the activated B cells enter the B-cell follicles, they proliferate in an environment provided by a network of FDCs and TFH cells. TFH cells were described in detail in Chapter 11, and both FDCs and TFHs provide survival signals to the rapidly dividing germinal center B cells. B-cell proliferation results initially in a decrease in IgD expression on the cell surface, although IgM expression is maintained. As the GC expands, nonresponding IgD⫹ B cells are displaced into the border region of the follicle, forming a corona of naïve B cells referred to as the follicular mantle zone. This is illustrated rather dramatically in Figure 12-11a as a blue border to the germinal center and is also illustrated in Figure 12-11b. Germinal center formation requires a number of cytokines, including Lymphotoxin-␣ (TNF-␤), that are produced by the FDCs, as well as by the TFH and potentially by other cells within the lymph node. These soluble signaling molecules help to repress genes such as blimp-1, and promote the expression of genes such as AID, which have important roles in the developing germinal centers. Dark- and Light-Zone Development As the germinal center matures, two zones become visible: the light zone and the dark zone (see Figure 12-11). The “darkness” of the dark zone results from a dense packing of rapidly proliferating B cells (centroblasts), whereas the “lighter” nature of the light zone results from the less dense distribution 398 PA R T V | Adaptive Immunity: Effector Responses of B cells (centrocytes) within a network of FDCs; indeed, FDCs are largely absent from the dark zone. (Note that in Figure 12-11a, not all the B cells have been labeled, in order to allow for clarity in the image. The green dark zone is, in fact, densely packed with dividing centroblasts, a characteristic that is better visualized in Figure 12-11b.) The distribution of B cells between the two zones is dependent on their relative levels of expression of the chemokine receptors CXCR4 and CXCR5, which interact with chemokines preferentially expressed in each of the two zones. CXCR4 is up-regulated on centroblasts, whereas CXCR5 has higher expression on centrocytes. Somatic Hypermutation and Affinity Selection Occur within the Germinal Center The term somatic hypermutation describes one of the most extraordinary processes in biology. The word somatic tells us that the mutational processes are occurring outside of the germ line (egg and sperm) cells. Hypermutation alludes to the fact that the mutational processes occur extremely rapidly. Somatic hypermutation (SHM) in mice and humans occurs only following antigen contact, affects only the variable regions of the antibody heavy and light chains, and requires the engagement of T cells, which must be able to interact with B cells via CD40-CD40L binding. The possibility that SHM of Ig genes might play a role in antibody diversification was first implied by amino acid sequencing studies conducted in the 1970s by Cesari, Weigert, Cohn, and others. Their studies focused on antibodies produced by mouse myeloma tumors expressing ␭ light chains. The mouse ␭ locus has been severely truncated, and as a result, mice have very few different ␭ chain variable regions. These investigators were therefore able to compare each of their myeloma light-chain sequences to a known germ-line sequence. They showed that the myeloma tumors expressed point mutations that were restricted to the variable regions of the ␭ chains and clustered in the complementarity determining regions. With the advent of nucleic acid sequencing technology, these data were confirmed and extended by others who showed that SHM affects the variable regions, but not the constant regions, of both the heavy and the light chains of Igs; that the frequency of somatic hypermutations increases with time post immunization and further; and that somatic hypermutation, followed by antigen selection, results in the secretion of antibodies whose affinity for the immunizing antigen increases with time post immunization. These experiments defined the process of SHM, but left open the question of where it occurred. A series of papers published in the early 1990s, and representing a technical tour de force, proved that somatic hypermutation occurs within the germinal centers of an active lymph node (see Classic Experiment Box 12-1). Once investigators knew where mutation was occurring, their next question, inevitably, was how? This question presented huge experimental challenges, and for almost 30 years investigators made only incremental progress in defining the mutational mechanism. Certain nucleotide sequences were defined that appeared to be particularly susceptible to mutation, and immunogeneticists were able to demonstrate that the mutational process appeared to be remarkably focused on the Ig variable regions. Other studies showed that SHM required the involvement of T cells, that it occurred only after antigen contact, and that mutation appeared to affect only genes that were being rapidly transcribed. However, the nature of the enzymes that mediate the response, and the molecular mechanisms involved, remained obscure. The breakthrough came, as so often happens, from an unexpected direction. In experiments designed to detect the genes and enzymes that mediate CSR, Takasu Honjo and colleagues isolated a cell line in which the antibody genes did not undergo CSR. This type of finding is often the crucial first step in identifying those proteins that are implicated in a particular process, as scientists could now compare those cells that could and could not undertake CSR and look for differences. However, importantly, they then noticed that the antibody genes in these cell lines were also devoid of somatic mutations in the variable regions. Could it be that the same enzyme was implicated in both SHM and CSR? Members of Honjo’s laboratory quickly isolated the gene that had been mutated in this cell line and proved that it encoded the enzyme Activation-Induced Cytidine Deaminase (AID), which was subsequently shown to be necessary for both somatic hypermutation and CSR. But now the scientists were faced with the new questions: How does AID induce hypermutation, and how is it restricted only to antibody variable regions? AID-Mediated Somatic Hypermutation Although all the details of the SHM mechanism are not yet completely understood, several parts of the process are now well established. AID-induced loss of the amino group from cytidine residues located in mutation hot spots results in the formation of uridine, as shown in Figure 12-12. This creates a U-G mismatch. Several alternative mechanisms then come into play that participate in the resolution of the original mismatch, leading to the creation of shorter or longer stretches of mutated DNA (Figure 12-13). NH2 O N O HN N O HO OH H Deoxycytidine O Activation induced cytidine deaminase N O HO OH H Deoxyuridine FIGURE 12-12 Activation-induced Cytidine deaminase (AID) mediates the deamination of cytidine and the formation of uridine. B-Cell Activation, Differentiation, and Memory Generation | CHAPTER 12 399 BOX 12-1 CLASSIC EXPERIMENT Experimental Proof That Somatic Hypermutation and AntigenInduced Selection Occurred within the Germinal Centers In a series of papers written in the early 1990s, the laboratory of Garnett Kelsoe conclusively demonstrated that the phenomena of somatic hypermutation and antigen-induced selection occurred within the germinal center regions of the secondary lymphoid organs. The researchers made use of the fact that the immune response to the T-dependent haptenprotein antigen—4-hydroxy-3-nitrophenylacetyl (NP)-hemocyanin—generates an immune response that uses a predictable combination of heavy- and light-chain variable regions. This enabled them to study the changes in antibody genes as the immune response progressed, whereas a similar study of a response that used, for example, a hundred or more different B-cell clones would have been technically impossible. The researchers generated serial tissue sections from lymph nodes at different time points during an immune response to the T-dependent antigen. The tissue samples were derived from the primary foci as well as from the germinal centers of antigen-responding lymph nodes. By scraping the primary foci or germinal center cells off the slides using microdissection tools, isolating the mRNA, and then subjecting the mRNA to RT polymerase chain reaction (PCR) and nucleic acid sequencing, mRNA from 100–300 B cells per sample were obtained. The results can be summarized as follows: • Each focus and germinal center is initiated by seeding with one to six antigenspecific B cells. Statistical analysis of the immunoglobulin variable region sequences obtained from each sam- ple showed that each focus or germinal center contained few B cells at the onset of the response. With time, the number of distinct clones of B cells per focus or germinal center decreased. • Somatic hypermutation is restricted to the germinal center B cells. Nucleic acid sequencing of immunoglobulin variable regions obtained from B cells in both primary foci and germinal centers demonstrated that foci-derived sequences did not show any evidence of mutation. In contrast, mutations could be detected in germinal centerderived, V-region genes and accumulated as the response progressed. • With time post immunization, mutations in the complementarity-determining regions are selected for; mutations in framework regions are selected against. Somatic hypermutation is an antigendriven process, unlike the other mechanisms for the generation of antibody diversity. As discussed in Chapter 7, the variable region of an immunoglobulin gene can be divided into framework and complementaritydetermining regions. Jacobs and colleagues analyzed what fraction of the mutations they observed were present in framework versus complementaritydetermining regions. Those mutations affecting framework residues might be expected to be either neutral with respect to antigen binding or, if they affected the folding of the antibody variable region, to be deleterious. In contrast, those in the complementaritydetermining regions of the gene, which encode the antigen-binding Briefly, the simplest mechanism is the interpretation of the deoxyuridine as a deoxythymidine by the DNA replication apparatus. In this case, one of the daughter cells would have an A-T pair instead of the original G-C pair found in the parent cell (Figure 12-13, left). Alternatively, the mismatched uridine could be excised by a DNA uridine glycosylase enzyme. Errorprone polymerases would then fill the gap as part of the cell’s region of the antibody molecule, have the potential to be advantageous. Jacob and Kelsoe asked whether selective mechanisms were at play that increased the frequency of B cells with enhanced antigen-binding capacity. Their results showed that, whereas at 8 days post immunization, mutations were randomly distributed throughout the Ig variable region gene, by the end of the primary immune response, a greater fraction of the mutations was located in the complementaritydetermining regions, despite the fact that these regions comprised only 21% of the V region sequence. • Affinity measurements of antibodies generated in vitro from PCR-amplified samples derived from foci and germinal center cells showed that mutated and selected antibodies displayed increased affinity for antigen. Thus, not only did Kelsoe’s lab demonstrate that the germinal center was the site of somatic hypermutation; their experiments also showed that mutations in the complementarity-determining regions that led to increased affinity for antigen were positively selected. Jacob, J., C. Miller, and G. Kelsoe. 1992. In situ studies of the antigen-driven somatic hypermutation of immunoglobulin genes. Immunology and Cell Biology 70:145–152. Jacob, J., and G. Kelsoe. 1992. In situ studies of the primary immune response to (4-hydroxy-3-nitrophenyl)acetyl. II. A common clonal origin for periarteriolar lymphoid sheath-associated foci and germinal centers. Journal of Experimental Medicine 176:679–687. short-patch base excision repair mechanism (Figure 12-13, middle). Third, mismatch repair (MMR) mechanisms could be invoked that result in the excision of a longer stretch of DNA surrounding the mismatch. The excised strand could then be repaired by error-prone DNA polymerases, such as DNA polymerase ␩, leading to a lengthier series of mutations in the region of the original mismatch (Figure 12-13, right). PA R T V 400 | Adaptive Immunity: Effector Responses C G AID U G Replication Uridine glycosylase U G MMR G Short-patch base excision repair Error prone polymerases A C T T A G G FIGURE 12-13 The generation of somatic cell mutations in Ig genes by AID. AID deaminates a cytidine residue, creating a uridine-guanosine (U-G) mismatch. Resolution of this mismatch may be facilitated by any one of several pathways that may compete with one another. On the left, the deoxyuridine is interpreted by the DNA replication machinery as if it were a deoxythymidine, resulting in the creation of an A-T pair in place of the original G-C pair in one of the daughter cells. In the center example, the mismatched uridine has been excised, most probably by one of the uridine DNA glycosylase enzymes, leaving an abasic site. The uridine can then be replaced by any of the four bases, in a reaction known as short-patch base excision repair, which can be catalyzed by one of a number of error-prone polymerases. Finally, as shown on the right, mismatch repair enzymes can detect the mismatch and excise a longer stretch of the DNA surrounding the U-G couple. Error-prone polymerases are then recruited to the hypermutable site by proliferating cell nuclear antigen (PCNA), and these polymerases can introduce a number of mutations around the original mismatch. Thus, depending on the repair mechanism, a mutation may occur only in the originally altered base or in one or more bases surrounding it. [Adapted from J. U. Peled et al., 2008, The Biochemistry of Somatic Hypermutation. Annual Reviews Immunology 26:481–511.] Mutational Apparatus Targeting Since the rate of hypermutation is orders of magnitude higher in Ig variable region DNA than in other genes in germinal center B cells, some mechanism must exist to direct the mutational machinery to the correct chromosomal location. In addition, a variation on this mechanism should also direct AID to those parts of the constant region genes that are recognized during CSR. When DNA is being actively transcribed, localized regions of DNA become transiently detached from their partner strands. The presence of single-stranded DNA appears to be necessary for SHM to occur, and the number of mutations that accumulate in Ig DNA is roughly proportional to the rate of Ig transcription. However, genes other than Igs are transcribed in germinal center B cells, as are the constant regions of the immunoglobulin genes, and they do not undergo mutation. Clearly this distinction alone does not account for the fact that mutations concentrate almost exclusively in the variable regions of Ig heavy and light chains, and something other than active transcription must be targeting AID to the variable regions. Careful analysis of Ig variable region sequences revealed that some sequence motifs were more likely than others to be targeted by the mutational apparatus, and these are referred to as mutational hot spots. In particular, it was noted that the sequence motif DGYW/WRCH was frequently targeted for mutation at the underlined G-C pair. (In this description of the sequence, the four nucleotides after the backslash represent the inverted complement of the DGYW sequence, such that the underlined G and C are paired with one another.) The code used to describe the targeted hot spot sequence is as follows: D ⫽ A/G/T Y ⫽ C/T R ⫽ A/G W ⫽ A/T H ⫽ T/C/A The DGYW motif is also found frequently in the class switch regions, and so it appears to be an important sequence that directs AID binding to certain parts of the DNA. Antigen-Induced Selection of B Cells with Higher Affinity In this section, we will describe how those B cells with higher-affinity antigen receptors successfully compete with their lower-affinity counterparts for survival and proliferation signals delivered to them by TFH cells in the germinal center. Seminal experiments using intravital fluorescence microscopy in transgenic mouse models have enabled investigators to address the question of how higher-affinity B cells are allowed to survive, while their lower-affinity precursors and cousins succumb to apoptosis. In these mice, antigen-specific T and B cells as well as FDCs were labeled with different fluorescent dyes, and the movement of B and T cells during an ongoing germinal center reaction was observed with timelapse fluorescence microscopy. Investigators observed that B cells spend a relatively short amount of time in the light zones of the germinal centers, where they must take the opportunity to come into contact with TFH cells, in order to receive T-cellderived growth and survival signals. They further noticed that there are many more B cells than TFH cells in the light zones, which indicated that B cells may have to compete with one another for the privilege of interacting with T cells. A well-accepted current model suggests that B cells bearing higher-affinity receptors capture and process antigen B-Cell Activation, Differentiation, and Memory Generation more effectively than do competing B cells with lower-affinity receptors. The more antigen that a B cell processes, the more antigen it will present on its class II MHC molecules for recognition by cognate TFH cells. Therefore, a B cell that has undergone an advantageous mutation in the antigenbinding region of its Ig genes will be better able to interact with TFH cells than will its competitors, and therefore it will receive more proliferative and survival signals from those TFH cells. Additional experiments have indicated that B-cell competition for antigen within the germinal center may be more direct than previously thought, as some B cells were observed actually stripping antigen from other B cells. This suggests that higher-affinity B cells actually steal antigen from their lower-affinity counterparts. In this way B cells with higheraffinity receptors present more antigen to T cells, and enjoy better interactions with them that lead directly to enhanced survival. In the absence of positive survival signals, loweraffinity B cells in the germinal center undergo apoptosis. Genetic dissection of this process has shown that signals delivered from the T cell through CD40 on the B-cell surface provide an indispensable component of the signal that T cells deliver to successfully competing B cells. In addition, BCRinduced PI3 kinase activation (see Chapter 3) in germinal center B cells results in the activation of the serine/threonine kinase Akt. Akt is a pleiotropic (has many effects) kinase, which not only promotes cell survival and inhibits apoptotic proteins, but also promotes the degradation of p53, thus allowing the GC B cells to cycle. Thus, those B cells that bind, process, and present more antigen to T cells will win out over those B cells that cannot express antigen as effectively. Low-affinity B cells will lose out in the competition for antigen, and will die because of the absence of T-cell-mediated survival signals. Since mutation is a random process, some B cells may acquire self-reactive receptors. One proposed mechanism for the destruction of such mutation-generated, self-specific B cells relies on the fact that self molecules, such as serum proteins, will be expressed at extremely high concentrations in the lymph nodes, whereas foreign antigens will be expressed at much lower levels. One might therefore expect that all the BCRs on a self-reactive B cell will be occupied. Full receptor occupancy leads to rapid internalization of the vast majority of the BCRs on the cell surface. In follicular B cells, such loss of cell surface BCR expression would result, in turn, in the loss of signaling through the BCR, exit from the cell cycle, and the induction of apoptosis. In contrast, on B cells specific for foreign antigens, only a relatively small proportion of their receptors will be occupied by antigen and therefore, on these cells, enough BCRs will remain on the cell surface to provide the scaffold for the signaling cascade. Inevitably, some self-specific B cells will escape into the periphery. Recall, however, that these B cells will lack cognate T-cell help and, in the absence of stimulation, will simply be lost by neglect. | CHAPTER 12 401 As described above, both SHM and CSR depend on the activity of the AID enzyme. However, analysis of mutated forms of AID has demonstrated that different parts of the same molecule catalyze the two different processes. Furthermore, mutations in different parts of the molecule lead to different immunodeficiency states. Specifically, investigators analyzed the structure of AID genes isolated from a series of patients with Hyper IgM syndrome. In this immunodeficiency disease, patients generate only IgM antibodies that fail to undergo either SHM or CSR. Such patients suffer from recurrent, severe infections, thus emphasizing the physiological importance of both the mutational and the class switching processes to the fully functioning immune system. However, some unusual patients were found to have AID genes with premature stop codons near the 3⬘ end. These patients generated antibodies that could undergo SHM, but were severely compromised in their ability to perform CSR. Such individuals displayed only mild symptoms, and their immunodeficiency was often not diagnosed until adulthood. This implies that different sections of the AID protein are necessary for SHM and for CSR and further suggests that the ability to generate high-affinity antibodies may be more functionally relevant than the capacity to synthesize antibodies of different heavy-chain classes. Class Switch Recombination Occurs Within the Germinal Center after Antigen Contact In Chapter 7, we noted that naïve B cells could simultaneously express both mIgM and mIgD: both proteins are encoded on the same long transcript, and the decision to translate ␮ (IgM) versus ␦ (IgD) heavy chains is made at the level of RNA splicing. In contrast, the decision to switch from the expression of IgM to expression of any of the other classes of antibodies is made at the level of DNA recombination, and the process by which it occurs is referred to as class switch recombination (CSR). The switch to the expression of any heavy-chain class other than ␦ results in the irreversible loss of the intervening DNA. The Ig heavy-chain locus is approximately 200 kb in length. The formation of ␥, ␧, and ␣ heavy-chain genes requires cutting and rejoining of the heavy-chain DNA (Figure 12-14) in such a way that the desired constant region lies directly downstream from the rearranged VDJ region. Class switching occurs by the induction of recombination between donor and acceptor switch (S) regions located 2 kb to 3 kb upstream from each CH region (except for C␦). The donor S region is the S region upstream from the antibody heavychain constant region gene expressed prior to the class switch (which is normally ␮, except for those instances in which a B cell undergoes more than one class switch). The acceptor S region is the S region upstream of the antibody heavy-chain constant region that the B cell will express next. Switch regions consist of tandem repeats of short, G-rich sequences, 20 bp to 80 bp in length, that differ slightly for PA R T V 402 | Adaptive Immunity: Effector Responses Heavy chain genes in IgM-expressing cells δ mRNA Germline transcript µ mRNA AID 5' V DJ AID Cµ Cδ C γ3 3' C γ1 S AID Makes double-strand breaks in S regions C γ 2b S S C γ 2a Cα Cε S S AID C γ1 C γ3 Class switch recombination (requires DSBs) C γ 2b Cδ C γ 2a Cµ + Heavy chain genes in IgE-expressing cell 5' V DJ Cα Cε 3' S FIGURE 12-14 Class switch recombination from a C␮ to a C␧ heavy-chain constant region gene. The activation induced cytidine deaminase (AID) enzyme initiates CSR by deaminating cytidine residues within the switch (S) regions upstream of C␮ and C␧ on both strands. This leads to the formation of double-strand breaks within both S regions that are then resolved by DNA repair mechanisms, with the loss of the intervening DNA sequence, as described in the text. [Adapted from J. Stavnezer et al., 2008, Mechanism and Regulation of Class Switch Recombination. Annual Review of Immunology 26:261–292.]. each isotype and contain targeting sites for AID. In CSR, genetic analysis has indicated that the critical sequence motif required for AID binding is a pair of WGCW overlapping motifs on the top and bottom strand, where W represents either adenine or thymine. (Note that WGCW merely represents a subset of the group of sequences described above as DGYW.) The overall length of the switch regions vary from 1 kb to 10 kb, and CSR can occur anywhere within or near the S regions. Signals for Class Switch Recombination B cells must receive costimulatory signals from CD40 or, occasionally, B-cell Toll-like receptors in order to engage in CSR. The importance of CD40-CD40L interactions in the mediation of CSR is illustrated in patients suffering from X-linked Hyper-IgM syndrome, an immunodeficiency disorder in which TH cells fail to express CD40L. Patients suffering from this disorder express IgM, but no other isotypes. Such patients also fail to form germinal centers, they fail to generate memory cell populations, and their antibodies do not show evidence of SHM. The cytokine signal received by the B cell determines which class of Ig it will make (see Table 12-1). These cytokines signal the B cells to induce transcription from germline promoters located upstream (5⬘) of the respective donor and acceptor switch regions. The resulting germ-line transcripts do not encode proteins and are therefore referred to as sterile RNAs. Importantly, no CSR can occur in the absence of this transcriptional activity, which is probably important in the creation of localized regions of singlestranded DNA recognized by AID. The germ-line promoters express the appropriate cytokine-responsive elements. For example, germ-line ␥1 and ␧ promoters, which are induced by IL-4, have binding sites for the IL-4-induced transcriptional activator Stat6. But how does a B cell that receives a signal from IL-4 know whether to switch to IgG1 or IgE? The answer to this, and similar questions, is not yet clear, and there is still much to learn regarding the details of the signals that differentially regulate switching to particular classes of antibodies. The Molecular Mechanism of Class Switch Recombination CSR occurs by an end-joining mechanism and, like SHM, the process is initiated by AID. AID deaminates several cytosines within both the donor and acceptor S sites that have been previously activated as a result of cytokine signaling. DNA uridine glycosylase enzymes remove the U, created by the deamination of cytidine, and then apurinic/apyrimidinic endonucleases nick the DNA backbone at the abasic sites, B-Cell Activation, Differentiation, and Memory Generation TABLE 12-1 Specific cytokines signal B cells to undergo CSR to different heavy-chain classes Cytokine signal Isotype synthesized by target B cell IL-4 IgG1, IgE TGF-␤ IgA, IgG2b IL-5 IgA IFN-␥ IgG3, IgG2a | CHAPTER 12 403 • What roles do residual antigen, follicular dendritic cells, and T-cell signaling play in the survival of B cells after the primary response is over? • Since some cells will survive this primary response as either memory B cells or long-lived plasma cells, how are B cells within a single clone selected for survival versus cell death? Is the decision made earlier or later during the course of the response, and is it random or directed in some as yet unknown way? Some Germinal Center Cells Complete Their Maturation as Plasma Cells creating single-strand breaks at multiple points in the donor and acceptor S sites. Mismatch DNA repair enzymes then convert the single-strand DNA breaks into double-strand breaks. In the final step of the process, the cell’s doublestrand break repair machinery steps in and ligates the two switch regions, resulting in the excision of the intervening sequence. The process of CSR can occur more than once during the lifetime of the cell. For example, an initial CSR event can switch the cell from making IgM to synthesizing IgG1, and a second CSR can switch it to making IgE or IgA. Most Newly Generated B Cells Are Lost at the End of the Primary Immune Response Between 14 and 18 days after its initiation, the primary immune response winds down, and the immune system is faced with a problem of excess. Rapid proliferation of antigenspecific B cells over the course of the immune response leads to the generation of expanded clones of cells, and if all the cells from each clone were allowed to survive at the close of every immune response, there would soon be no room for new B cells emerging from the bone marrow and seeking to circulate through the lymphoid follicles and receive their survival signals. Although we know that most B cells are lost by apoptosis at the end of the immune response, the exact mechanism by which this occurs has not yet been fully characterized. As antigen levels wane, the balance between survival signals and death signals experienced by the B cell in the lymph node may be tipped in favor of apoptosis. Certainly, the B cell no longer receives survival signals via antigen binding to the BCR. In addition, animals deficient in the genes encoding Fas have excess numbers of B cells, implicating the Fas-FasL interaction in the control of B-cell numbers. But we do not yet know the whole story. This is an area of active research, and a number of important questions remain to be answered: • Does the cessation of B-cell signaling simply lead to a decrease in anti-apoptotic proteins, such that the cell is no longer protected from these death-inducing signals? Or is a more active switch engaged that leads to the cell’s demise? At some point in the ongoing immune response, approximately 5 to 15 days after antigen encounter, a fraction of germinal center B cells will begin to up-regulate IRF-4 expression, heralding the beginning of their differentiation into antibody-secreting plasma cells. As described earlier, IRF-4 expression then induces the generation of the transcriptional repressor, BLIMP-1, which down-regulates those genes important to B-cell proliferation, CSR, and SHM, and up-regulates the rate of the synthesis and secretion of Ig genes. As the germinal center B cell differentiates into a fully mature plasma cell, it reduces the level of expression of the chemokine receptor CXCR5, which has been responsible for retaining it within the germinal center. Instead the nascent plasma cell begins to express CXCR4, which enables it to leave the lymph node and circulate within the peripheral tissues. These germinal center-derived plasma cells differ from those generated from the primary focus in that their Ig genes have undergone both SHM and CSR, and hence the antibodies they secrete will be of high affinity and may be class switched. For many years, it was thought that plasma cells localized primarily in the medullary cords of the lymph nodes (the inner parts of the kidney-shaped lymph nodes), or the red pulp in the spleen, and that they were relatively short-lived. However in the past decade or so, we have come to understand that plasma cells can home to several other locations and that the 10% to 20% of plasma cells that home to the bone marrow can be very long-lived. Indeed, smallpox-specific serum antibodies have been identified 75 or more years after immunization with smallpox vaccine, suggesting that the plasma cells secreting these antibodies may persist for the lifetime of the host. We now know that these long-lived plasma cells derive from both the plasma cells of the primary focus as well as from the B cells that have passed through the germinal center and have undergone CSR and SHM. Within the bone marrow, the niches occupied by fully differentiated plasma cells differ from those inhabited by developing B cells (Figure 12-15). Experiments using in vitro culture techniques to determine the survival requirements for long-lived bone marrow plasma cells have highlighted the need for CXCL12 (recognized by CXCR4 on the plasma cell) and the TNF family cytokine member APRIL (recognized 404 PA R T V | Adaptive Immunity: Effector Responses Periphery Germinal center Plasmablast High-affinity antibodies B cell CXCR4 APRIL Plasma cell (CXCR4+) Eosinophil (CXCR4+) BCMA Megakaryocyte Stromal cells (CXCL12+) FIGURE 12-15 The bone marrow niche occupied by plasma cells is supported by eosinophils and megakaryocytes, as well as by mesenchymal stromal cells. Plasmablasts and plasma cells that have passed through the germinal center reaction and entered the circulation take up residence in the bone marrow, where they seek niches adjacent to eosinophils and by the plasma cell APRIL receptor, BCMA). CXCL12 is produced in the bone marrow by mesenchymal stromal cells, and APRIL by both eosinophils and megakaryocytes. Characterization of the bone marrow plasma cell niche continues to be a rapidly advancing area of research. Other plasma cells, generated in the lymphoid tissues of the gut, remain associated with the mucosal tissues and secrete large amounts of IgA. IgA-secreting plasma cells are generated both in the Peyer’s patches, areas of lymphoid concentration within the gut tissues, or in isolated lymphoid follicles in the lamina propria of the gut (see Figures 2-11 and 2-12). These gut-associated, IgA-secreting plasma cells have significant functional differences from the IgG-secreting plasma cells in other lymphoid tissues, and share some important characteristics with cells of the granulocytemonocyte lineages. In particular, they produce the antimicrobial mediators TNF-␣ and inducible nitric oxide synthase (iNOS), molecules normally associated with monocyte and granulocyte activation. In order to continue producing these Bone marrow megakaryocytes, as well as to the traditional mesenchymally derived stromal cells. The eosinophils and megakaryocytes provide the longlived plasma cell with the survival factor APRIL, a TNF family member recognized by the receptor BCMA, whereas the stromal cells release CXCL12, recognized by the receptor CXCR4 on plasma cells. mediators, gut IgA-producing plasma cells must remain in contact with gut stromal cells and be subject to microbial costimulation. Deletion of TNF-␣ and iNOS in B-lineage cells resulted in poor clearance of gut pathogens and has been associated with a concomitant reduction in IgA synthesis. B-Cell Memory Provides a Rapid and Strong Response to Secondary Infection The first recorded concept of immunological memory appears in the writings of Thucydides, around 430 bce. Describing the plague in Athens he wrote, “Only those who had recovered from the plague could nurse the sick, because they would not contract the disease.” Implicit in this statement was the fact that those who had suffered and recovered had immunity to the plague, and were therefore able to mount a stronger and faster response to future infection. Figure 12-16 illustrates the classic conception of a memory immune response. The index of immune responsiveness B-Cell Activation, Differentiation, and Memory Generation Antibody concentration in serum, units per ml 100 Total 10 Primary response Secondary response 1.0 Total IgG IgM 0.1 IgG IgM lag 0.01 1° Ag 2° Ag Time after immunization FIGURE 12-16 Concentration and isotype of serum antibody following primary and secondary immunization with antigen. The antibody concentrations are plotted on a logarithmic scale. The time units are not specified because the kinetics differ somewhat with type of antigen, administration route, presence or absence of adjuvant, and the species or strain of animal. in this figure is the production of serum antibody. The primary response is characterized by a lag period, which reflects the time required for the division and differentiation of B cells within the primary foci and their movement into the germinal center. The B cells of the primary foci then release IgM, and after a short delay B cells that have migrated into the germinal center join the response with the concomitant release of both IgM and IgG. As the primary response draws to a close, somatically hypermutated receptors make their first appearance, and then a selected subpopulation of B cells leaves the germinal centers and enters the memory B-cell compartment. The secondary response to the antigen is both faster and stronger than the first. Antigen-specific B-cell division has already occurred, and an expanded set of memory B cells bearing high-affinity receptors is available for immediate differentiation to high-affinity IgG secretion. Interestingly, somatically TABLE 12-2 | CHAPTER 12 405 hypermutated B cells can undergo further hypermutation with additional antigen exposure, such that the average affinity of antigen-specific antibodies increases through the third, and even a fourth, immunization with the same antigen. The actual kinetics of both primary and secondary B-cell responses varies with the nature, dose, and route of administration of the antigen. For example, immunization of mice with an antigen such as sheep red blood cells typically results in a lag phase of 4 to 5 days before antibody is reliably detected in serum, and peak serum antibody levels are attained by 7 to 10 days. In contrast, the lag phase for soluble protein antigens is a little longer, often lasting about a week, and peak serum titers do not occur until around 14 days. What properties distinguish memory from naïve B cells (Table 12-2)? Memory and naïve B cells share the characteristics of being nonsecreting cells that need further antigen stimulation prior to expansion and differentiation to plasma cells. However, memory B cells differ from their naïve counterparts in terms of their sensitivity to antigen activation; memory B cells are able to respond to lower concentrations of antigen than primary naïve B cells can, as a consequence of having generated higher-affinity Ig receptors during the process of SHM. They also respond more quickly to antigen activation, and therefore there is a shorter lag between antigen contact and antibody synthesis. Memory B cells do not express IgD and usually carry Ig receptors that are class switched (although some IgM-bearing memory cells may persist). Memory B cells have longer half-lives than primary B cells and recirculate among the lymphoid tissues. Despite the fact that scientists have known about the existence of these cells for decades, surprisingly little is known about the factors that control their generation and maintenance. Questions regarding memory B cells that still await resolution include the following: • At what stage in B-cell differentiation is the B-cell fate determined, and what are the signals that control that fate? It seems clear that at some point in the maturation process of a B cell, individual daughter cells of the same stimulated clone of primary B cells will be selected for memory generation (life) or death. What mechanism Functional differences between primary and secondary B cells Naïve B cell Memory B cell Lag period after antigen administration 4–7 days (depends on antigen) 1–3 days (depends on antigen) Time of peak response 7–10 days 3–5 days Magnitude of peak antibody response Varies, depending on antigen Generally 10 to 1000 times higher than primary response Antibody isotype produced IgM predominates in early primary response IgG predominates (IgA in the mucosal tissues) Antigens Thymus independent and thymus dependent Primarily thymus dependent Antibody affinity Low High Life span of cells Short-lived (days to weeks) Long-lived, up to life span of animal host Recirculation Yes Yes 406 PA R T V | Adaptive Immunity: Effector Responses allows for the differential survival of members of the same clone of B cells? Exciting, ongoing experiments that study asymmetric cell division in antigen-stimulated B cells are beginning to unravel the threads of this particular immunological knot. • Where, in the body, do memory cells bearing differentially class-switched receptors reside? In mice, some memory B cells localize to the marginal zones of the spleen, and others have been located in the follicles; in humans, memory cells appear to localize to the marginal zones and to sub- and intraepithelial surfaces. Interestingly, during an ongoing infection or in an individual suffering from an autoimmune disease, ectopic reservoirs of organized lymphoid tissues can develop in multiple organs not usually associated with immune responsiveness, including the kidneys, pancreas, lungs, thyroid, and so on. The contributions of such ectopic tissues, as well as of organized lymphoid organs within the bone marrow, gut, and other mucosal tissues, to memory responses have yet to be investigated. • How does the balance of pro- and anti-apoptotic molecules differ between primary and memory B cells, and are particular cytokine signals required for persistent memory B-cell survival? • Are there differences in the efficiency of signal transduction in the two kinds of cells that may help to explain the variations in their lag times to antibody synthesis? T-Independent B-Cell Responses Not all antibody-producing responses require the participation of T cells, and certain subsets of B cells have evolved mechanisms so as to respond with antibody production to particular classes of antigens, without T-cell help (see Figure 12-3). Antigens capable of eliciting T-independent antibody responses tend to have polyvalent, repeating determinants that are shared among many microbial species. Many of these antigens are recognized by B cells of the B-1 as well as the marginal zone subsets (discussed below). B-1 B cells secrete mainly IgM antibodies that are not subject to SHM. Because of the shared nature of their antigens and their oligoclonal antibody response, B-1 B cells may be considered to belong to an “innate-like” category of lymphocytes. Our understanding of the physiology of MZ is still developing, but it is clear that B cells of this unusual subset of cells may receive help from cell types other than T cells, as is discussed below (see Advances Box 12-2). T-Independent Antigens Stimulate Antibody Production without the Need for T-Cell Help The nude (nu/nu) mouse is one of the most bizarre accidents of nature. Devoid of body hair, its ears seem oversized, and it appears absurdly vulnerable (Figure 12-17). These mice have FIGURE 12-17 The nude mouse. Nude mice have a mutation in the Foxn1 gene that results in hair loss (alopecia) and interferes with thymic development, such that they possess very few T cells and only a thymic rudiment. Nude mice provided useful early models for the exploration of a T-cell-depleted immune response. [David A. Nortcott/Corbis] a mutation in the gene for the transcription factor Foxn1 that, in addition to affecting hair growth, also results in athymia: mice and humans with this mutation have only thymic rudiments and possess few mature recirculating T cells. Using nude mice as well as mice whose thymuses were surgically removed early in life (referred to as neonatally thymectomized), immunologists showed that most protein antigens fail to elicit an antibody response in such animals even while the response to many carbohydrate antigens was unaffected. Those antigens capable of generating antibody responses in athymic mice were referred to as T-independent antigens because they did not require T-cell help to generate an antibody response. Since T-cell interactions are required for the induction of AID, TI antigens elicit predominantly low-affinity antibody responses from B cells expressing only the IgM isotype. We now know that T-independent antigens fall into two subclasses (Table 12-3). TI-1 Antigens The first class of T-independent antigens (TI-1 antigens) is exemplified by the bacterial lipopolysaccharide (LPS). TI-1 antigens bind to innate immune receptors on the surface of all B cells (including the majority B-2 B-cell population), and are capable, at high-antigen doses, of being mitogenic for all B cells bearing the responding innate receptors. Since B-cell stimulation in this instance is occurring through the innate receptor (TLR4), only a small minority of the antibodies produced will be able to bind directly to the TI-1 antigen. The huge in vivo polyclonal TI-1 responses generated in response to high levels of Gram-negative organisms can be catastrophic for an individual and are associated with the phenomenon we know as septic shock (see Chapter 15). At lower doses, the innate immune receptors are unable to bind sufficient antigen to be stimulatory for the B cell. However, B-Cell Activation, Differentiation, and Memory Generation TABLE 12-3 | CHAPTER 12 407 Properties of thymus-dependent and thymus-independent antigens TI antigens Property TD antigens Type 1 Type 2 Chemical nature Soluble protein Bacterial cell-wall components (e.g., LPS) Polymeric protein antigens; capsular polysaccharides Isotype switching Yes No Limited Affinity maturation Yes No No Immunologic memory Yes No No Polyclonal activation No Yes (high doses) No Humoral response in those B cells that bind to the TI-1 antigen through their Ig receptors, the TI-1 antigen may still be able to cross-link the Ig and innate receptors, thereby eliciting an oligoclonal (few clones) B-cell response that remains independent of T-cell involvement. Under these circumstances, all the secreted antibodies will be specific for the TI-1 antigen. Given the number of commensal bacteria in the gut, many of which bear these TI-1 antigens, why aren’t our gut B cells in a constant state of activation? Part of the answer to this question lies in the anatomy of the gut. The gut is lined by a mucous layer, which prevents access of most commensal bacteria to gut-resident lymphocytes. Second, regulatory T cells in the gut help to tone down the immune response, thus preventing the inflammation that would ensue were organisms to mount a constant immune response to commensal organisms. TI-2 Antigens Unlike TI-1 antigens, type 2 T-independent antigens (TI-2 antigens), such as capsular bacterial polysaccharides or polymeric flagellin, are not mitogenic at high concentrations. Their capacity to activate B cells in the absence of T-cell help results from their ability to present antigenic determinants in a flexible and remarkably multivalent array, which causes extensive cross-linking of the BCR. In addition, most naturally occurring TI-2 antigens are characterized by the ability to bind to the complement fragments, C3d and C3dg (see Chapter 6). As a result, TI-2 antigens are capable of activating B cells by cross-linking the BCR and CD21 receptors on the B-cell surface. TI-2 antigens can only partially stimulate B cells in the complete absence of help from other cells. Monocytes, macrophages, and dendritic cells have been shown to facilitate B-cell responses to TI-2 antigens by expressing a molecule known as BAFF—a membrane-bound homolog of tumor necrosis factor—to which mature B cells bind though the TACI (transmembrane activator and CAML interactor) receptor. This interaction activates important transcription factors that promote B-cell survival, maturation, and antibody secretion. T cells can also enhance B-cell activation by producing cytokines that push TI-2 antigen-activated B cells from activation to antibody production. Recognition by B-cell-bound receptors of these T-cell cytokines may stimulate the B cell to secrete antibody classes other than IgM in response to TI-2 antigen stimulation. Unlike TI-1 antigens, TI-2 antigens cannot stimulate immature B cells and do not act as polyclonal activators. Two Novel Subclasses of B Cells Mediate the Response to T-Independent Antigens All B cells bear Ig receptors and secrete antibodies, but recent research has demonstrated that there are multiple subpopulations of B cells varying in locations, phenotypes, and functions (Table 12-4). Some of these subpopulations (the transitional B-cell populations, T1 and T2) represent temporal stages of B-cell development that occur after the B cell leaves the bone marrow. An additional transitional subset, T3, appears to represent an anergic subpopulation of B cells (see Chapter 10). Others (B-1a, B-1b, B-2, and marginal zone B cells) represent different subpopulations of mature B cells, each characterized by a preferential location and range of functional capacities. The T-dependent B-cell response described above is conducted by the B-2 cell population, also known as follicular B cells. The presence of B cells with properties distinct from those of the majority B-2 subset was first suggested when scientists noticed that B cells responding to T-independent antigens differed in several important ways from B cells recognizing T-dependent antigens. Those B cells with specificity for T-independent antigens can be divided into two major subtypes that differ in aspects of their development, their anatomical location, and their cell surface markers. We will describe each of them in turn. B-1 B Cells Like ␥␦ intraepithelial T cells, B-1 B cells occupy a functional niche midway between those of the innate and adaptive 408 PA R T V | Adaptive Immunity: Effector Responses ADVANCES New Ideas on B-Cell Help: Not All Cells That Help B Cells Make Antibodies Are T Cells numbers of neutrophils (neutropenic individuals) or with other neutrophil disorders. Such patients show reduced levels of IgM, IgG, and IgA antibodies to microbial TI antigens normally recognized by MZ B cells, whereas the amounts of these Igs with specificity for T-dependent (TD) antigens were normal. Furthermore, in cases where somatic hypermutation could be assessed, patients with neutrophil disorders showed less evidence of V-region mutations in MZ B cells than did healthy control subjects. These findings all suggest that neutrophils may play an important role in the survival and activation of MZ B cells in vivo, a suggestion borne out by the careful microscopic and in vitro culture experiments described by Puga et al. Neutrophils were observed colonizing the regions of the spleen around the MZ, even in the absence of infection or inflammation. Although microscopic observation and flow cytometry failed to detect major morphological distinctions between the MZ splenic neutrophils and immune systems; many representatives of both cell types are located outside the classical secondary lymphoid tissues, and the repertoires of their antigen receptors are less diverse than those of other, more conventional adaptive immune lymphocyte subsets. Both ␥␦ intraepithelial T cells and B-1 B cells are self-renewing in the periphery—their populations are maintained without the need to be continuously reseeded from bone-marrow-derived precursors. Finally, both of these subsets respond rapidly to antigen challenge with relatively low-affinity responses. The existence of the B-1 B-cell subpopulation was first described in 1983, when the lab of Leonard and Leonore Herzenberg discovered a set of B cells bearing the CD5 antigen, whose expression had previously been thought to be restricted to T cells. These CD5⫹ B cells were termed B-1 B cells to distinguish them from the conventional B-2 B-cell population. (The numbering reflects the order of appearance of the two B-cell subsets in ontogeny.) In humans and mice, B-1 B cells make up only about 5% of B cells, although in some species such as rabbits and cattle, those in the circulation, subsequent functional assays showed that the splenic neutrophils were able to induce IgM secretion by splenic MZ cells, even in the absence of prior neutrophil activation by antigen. Surprisingly, the splenic neutrophils were more effective in the induction of IgM secretion by MZ B cells than CD4⫹ T cells (Figure 1). Culture of the 1,200 IgM (ng/ml) Since the classical experiments of Miller, Mitchell, Mitchison, and others in the 1960s, immunologists investigating the sources of help for B-cell activation and antibody production have been conditioned to think in terms of T lymphocytes. However, we are gradually coming to appreciate that cells other than T lymphocytes may promote B-cell activation and survival. We now know that mesenchymal stromal cells and eosinophils cooperate in ensuring plasma cell survival in the bone marrow. Recently, in what can be considered a major shift in our thinking about aiding B-cell function, Puga and colleagues have demonstrated that marginal zone (MZ) B cells may receive help in activation, somatic hypermutation (SHM), class-switch recombination (CSR), and antibody secretion from an unexpected source: neutrophils. Since many cells are capable of exerting effects in vitro that are not replicated when tested in vivo, we first note that the numbers and/or activity of MZ B cells are diminished in patients with reduced 600 0 NBH DCS MS TS FIGURE 1 NBHs enhance IgM secretion by MZ B cells even more effectively than do helper T cells. Splenic MZ B cells were cultured for 6 days with NBH cells, splenic dendritic cells (DCs), macrophages (Ms), or CD4⫹ T cells (Ts). The concentration of IgM released into the culture supernatant was assessed by enzyme-linked immunosorbent assay (ELISA, see Chapter 20). B-1-like cells represent the major subset. However, even in humans and mice, B-1 cells predominate in the pleural and peritoneal cavities, and it is probable that their major function is to protect these body cavities from bacterial infection. Because of their priority in the B-cell developmental sequence, B-1 B cells are found in relatively high numbers in fetal and neonatal life. Cells having the functional characteristics of B-1 cells but lacking expression of the CD5 molecule were identified at a later stage and termed B-1b B cells. Because the lack of T-cell involvement in their stimulation means that AID is never activated, B-1 B cells secrete antibodies of relatively low affinity that are primarily of the IgM class. Since B-1 B cells derive from a limited number of B-cell clones generated early in ontogeny, the antibodies they secrete are also significantly less diverse than the antibodies secreted by B-2 B cells. Antibodies secreted by B-1 B cells have evolved to recognize antigenic determinants expressed by gut and respiratory system bacteria and are primarily directed toward such common repeated antigens as phosphatidyl choline (a component of pneumococcal cell walls), lipopolysaccharide, B-Cell Activation, Differentiation, and Memory Generation | CHAPTER 12 409 BOX 12-2 of soluble factors, or does cell-cell contact also play a role? The answer to this question appears to be both. Comparing the mRNA and protein expression, as well as the cell surface presence of a variety of helper B-cell factors between the circulating, inactive neutrophils (NC) and the NBHs, Puga et al. showed that the helper B-cell neutrophils produced higher quantities of soluble BAFF and APRIL, as well as of IL-21 than did the NC * 50 Relative amount of AID mRNA splenic neutrophils, termed B helper neutrophils, or NBH by the authors, with CD4⫹ T cells resulted in suppression of T-cell proliferation following stimulation with anti-CD3 and IL-2. This shows that the neutrophils are not stimulating B cells indirectly by first activating T cells, and also allows for the possibility that the neutrophils may be selectively biasing splenic B cells toward T-independent responses. Using qPCR, Puga and colleagues demonstrated the presence of aicda mRNA (encoding the AID protein) in MZ B cells. They further showed that culturing B cells in medium conditioned by the growth of NBH cells in it (NBH-conditioned medium) resulted in the up-regulation of aicda mRNA expression in those B cells. The amount of aicda mRNA in the MZ B cells was greater than in naïve B cells, but not as much as in B-2 B cells engaged in an active germinal center reaction. Culturing MZ B cells with NBH cells, or with NBH-conditioned medium, resulted in both SHM and CSR and confirmed that NBH-derived help was sufficient to drive both processes in the MZ B cells (Figure 2). What form does this neutrophilderived help take? Is it purely in the form 25 0 Ctrl NBH FIGURE 2 Quantitative RT PCR analysis of the expression of the aicda gene in MZ B cells cultured for 2 days in control medium (left) or in medium conditioned by NBH cells (right). The cells. They also expressed more BAFF on their membranes. Investigators reduced the amount of each of these cytokines that was available to B cells by adding soluble forms of the cytokine receptors to in vitro B cell cultures. The soluble receptors act by blocking B cell recognition. Under these conditions, IgM secretion, and CSR to IgG and IgA were both significantly reduced, thus demonstrating the relevance of the three cytokine signals to MZ B cell activation. The authors also note that NBH cells express surface CD40L, although at lower levels than those found on CD4+ T cells. Thus, at least for MZ B cells, the role played by helper neutrophils in B-cell activation may be even more relevant than that played by helper T cells. Paradigm shifts such as this one enliven a field in delightful and stimulating ways. Source: I. Puga, et al. 2011. B cell-helper neutrophils stimulate the diversification and production of immunoglobulin in the marginal zone of the spleen. Nature Immunology 13:170–180. relative amounts of mRNA encoding the AID protein were assessed by qPCR. Gene expression was normalized to the expression of pax5 mRNA. Culture of MZ B cells in NBH⫺ conditioned medium was shown to enhance the expression of the aicda gene by approximately 35-fold. and influenza viruses. B-1 B cells represent the majority component of the B-cell responses to TI antigens. Very early in the history of immunology, investigators noted that the serum of unimmunized mice and humans contains so-called natural IgM antibodies that bind a broad spectrum of antigens with relatively low affinity. These antibodies derive mainly from B-1 B cells, and their presence in unimmunized animals suggests that B-1 B cells may exist in a partially activated state and constitutively secrete low levels of natural antibodies. In addition, a relatively high frequency of these IgM antibodies display autoimmune reactivities, although their affinities for self antigens are sufficiently low that they rarely induce disease. Low-level interactions with self antigens during development may therefore be important in the development of B-1 B-cell function and in the maintenance of their partially activated phenotype. Although antibody secretion by B-1 B cells is not dependent on T-cell help, it can be enhanced in the presence of T-cell cytokines. Indeed, recent data suggests that, in the presence of T-cell help, B-1b cells may express certain attri- butes of B-2 cells, such as Ig class switching (with the resultant production of IgA antibodies), SHM, and the generation of a long-lasting antibody response. Marginal Zone B Cells The second class of B cells capable of responding to TI antigens is the marginal zone (MZ) subset, which resides in the marginal zone of the spleen (Figure 12-18). As for B-1 B cells, maintenance of physiological levels of MZ B cells appears to depend on their capacity to receive lowlevel signals through the BCR. Again, like B-1 B cells, MZ B cells have the capacity for self-renewal in the periphery and do not need to be constantly replenished from bone marrow precursors. MZ B cells derive originally from the transitional T2 B-cell population, and it has recently been shown that the Notch2 signaling system plays a role in sending B cells down the MZ pathway. This is notable because Notch signaling was previously thought to be restricted in its activity among lymphocytes to T cells. MZ B cells bear unusually high levels of CD21 (CR2), enabling 410 TABLE 12-4 PA R T V | Adaptive Immunity: Effector Responses Functional differences among mature B-cell subsets Attribute Conventional B2 B cells B-1 B cells Marginal zone B cells Major sites Secondary lymphoid organs Pleural and peritoneal cavity; also spleen Marginal zones of spleen in mice; primates also have MZ cells in other locations V region diversity Highly diverse More restricted diversity Moderate diversity Rapidity of antibody response Slow; plasma cells appear 7–10 days post stimulation Rapid; plasma cells appear as early as 3 days after stimulation Rapid; plasma cells appear as early as 3 days after stimulation Surface IgD? High levels of IgD Low levels of IgD Low levels of IgD Somatic hypermutation Yes No Yes in primates; possibly in rodents Requirements for help from other cell types Provided by T cells No, although T and other cells can enhance response Dendritic cells and neutrophils can enhance response (see Box 12-2) Participate in germinal center reaction? Yes No Possibly, although with slower kinetics than follicular B cells Isotypes produced All isotypes Predominantly IgM Predominantly IgM Immunological memory Yes Very little Yes; source of IgM-producing memory cells them to bind very efficiently to antigens covalently conjugated to C3d or C3dg. Since one characteristic of TI-2 antigens is an enhanced tendency to bind C3d, MZ B cells are particularly important in the host protection against pathogens bearing TI-2 antigens. MZ B cells are specialized to respond to blood-borne antigens that enter the immune system via the splenic MZ. Antigen stimulation of MZ B cells results in their movement from the MZ to the bridging channels and red pulp of the spleen, where they undergo a burst of proliferation, forming foci of plasmablasts not unlike the primary foci formed in the lymph Follicle Marginal sinus Follicular arteriole nodes upon antigen challenge. These cells produce high levels of antigen-specific IgM within 3 to 4 days after antigenic stimulation. Very recent work has suggested that MZ B cells may be helped in antibody secretion, SHM, and CSR not by T cells (as we have come to expect), but instead by neutrophils, which we normally think of as participants in the innate arm, rather than the adaptive arm, of immunity. These experiments are described in Advances Box 12-2. Some of the characteristics of MZ B cells, as well as their local environments, differ between rodents and primates. Of particular interest is the possibility that the variable regions Marginal zone Periarteriolar lymphoid sheath Central arteriole Venous drainage Arterial supply FIGURE 12-18 Longitudinal section through the mouse spleen, showing its blood supply and the location of lymphoid populations. The periarteriolar lymphoid sheath (PALS), containing the T-cell zone, can be seen surrounding the central arteriole, outside of which are the lymphoid follicles. Outside the follicles lies the marginal sinus, and the marginal zone separates the sinus from the red pulp of the spleen. The anatomy of the marginal regions varies somewhat between primate and rodent spleens. [Adapted from S. Pillai et al., 2005, Marginal Zone B Cells. Annual Review of Immunology 23:161–196, Figure 1.] B-Cell Activation, Differentiation, and Memory Generation of antibodies derived from human MZ B cells may undergo SHM in the absence of an obvious germinal center reaction, and conceivably even in the absence of antigenic stimulation. Furthermore, in rodents, B cells with the characteristics of MZ cells are restricted to the MZs of the spleen, whereas in primates they can be found in other peripheral lymphoid tissues such as the tonsils. Negative Regulation of B Cells Up until this point, we have been discussing how B cells are activated, and the functional correlates of that activation. However, antigen stimulation of B cells results in a proliferative response that is as rapid as any observed in vertebrate organisms; an activated lymphocyte may divide once every 6 hours. Control mechanisms have therefore evolved to ensure that B-cell proliferation is slowed down once sufficient specific B cells have been generated and that most of the B cells enter into an apoptotic program once the pathogen has been eliminated. In this section, we will address negative regulation of B-cell activation that is mediated through two different molecules on the B-cell surface. Negative Signaling through CD22 Shuts Down Unnecessary BCR Signaling In addition to CD19/CD21 and CD81, the BCR of resting B cells is also associated with an additional transmembrane molecule, CD22. CD22 bears an Immunoreceptor Tyrosinebased Inhibitory Motif (ITIM), similar in structure to the ITAM motifs introduced in Chapter 3, but mediating inhibitory, rather than activating functions. Activation of B cells results in phosphorylation of the ITIM, thus allowing association of the SHP-1 tyrosine phosphatase with the cytoplasmic tail of CD22. SHP-1 can then strip activating phosphates from the tyrosine of neighboring signaling complexes. For as long as the BCR signaling pathway is being activated by antigen engagement, phosphate groups are reattached to the tyrosine residues of adapter molecules and other signaling intermediates as fast as the phosphatases can strip them off. However, once antigen levels begin to decrease, receptor-associated tyrosine kinase activity slows down, and signaling through CD22 can then induce the removal of any residual activating phosphates. CD22 thus functions as a negative regulator of B-cell activation, and its presence and activity ensure that signaling from the BCR is shut down when antigen is no longer bound to the BCR. Consistent with this negative feedback role, levels of B-cell activation are elevated in CD22 knockout mice, and aging CD22 knockout animals have increased levels of autoimmunity. CD22 is a cell-surface receptor molecule that recognizes N-glycolyl neuraminic acid residues on serum glycoproteins and other cell surfaces and can thus double as an adhesion molecule. It is expressed in mature B cells that bear both mIgM and mIgD Ig receptors. | CHAPTER 12 411 Negative Signaling through the Fc␥RIIb Receptor Inhibits B-Cell Activation It has long been known that the presence of circulating, specific, antigen-IgG complexes is inhibitory for further B-cell activation, and this phenomenon has now been explained at the molecular level by the characterization of the Fc␥RIIb receptor (also known as CD32). Fc␥RIIb recognizes immune complexes containing IgG and, like CD22, bears a cytoplasmic ITIM domain. Co-ligation of the B cell’s Fc␥RIIb receptor molecules with the BCR by a specific antigen-antibody immune complex results in activation of the Fc␥RIIb signaling cascade, and phosphorylation of Fc␥RIIb’s ITIM. The phosphorylated ITIM serves as a docking site for the inositol phosphatase SHIP, which binds to the ITIM via its SH2 domain. SHIP then hydrolyzes PIP3 to PIP2, thus interfering with the membrane localization of the important signaling molecules Btk and PLC␥2 (see Chapter 3) and causing the effective abrogation of B-cell signaling. Signaling through Fc␥RIIb also results in decreased phosphorylation of CD19 and reduced recruitment of PI3 kinase to the membrane. Negative signaling through the Fc␥RIIb receptor makes intuitive sense, as the presence of immune complexes containing the antigen for which a B cell is specific signals the presence of high levels of antigen-specific antibody, and hence a reduced need for further B-cell differentiation. B-10 B Cells Act as Negative Regulators by Secreting IL-10 Recently, an unusual population of B cells has been discovered that appears to be capable of negatively regulating potentially inflammatory immune responses by secreting the cytokine IL-10 upon stimulation. Working with two mouse autoimmune models, investigators demonstrated that B cells capable of secreting the immunoregulatory cytokine IL-10 could alleviate the symptoms of a mouse suffering from a form of the antibodymediated autoimmune disease multiple sclerosis. Recall from Chapter 11 that IL-10 is a cytokine normally associated with regulatory T cells. It has pleiotropic effects on other immune system cells, which include the suppression of T-cell production of the cytokines IL-2, IL-5, and TNF-␣. Furthermore, IL-10 interacts with antigen-presenting cells in such a manner as to reduce the cell surface expression of MHC antigens. The finding that B cells could be capable of secreting this immunoregulatory cytokine represents the first indication that they, as distinct from T cells, might have the capacity to down-regulate the function of other immune system cells. A small population of splenic B cells appears to account for almost all of the B-cell-derived IL-10. However, at this point, we do not know whether this IL-10 secreting B-cell population truly represents a single developmental B-cell lineage. For example, B cells producing IL-10 have been identified among both B-1 and B-2 412 PA R T V | Adaptive Immunity: Effector Responses B-cell populations. In addition, some, but not all B cells producing IL-10 bear markers typical of the transitional T2 B-cell subset. Importantly, all B-10 B cells appear to demonstrate the capacity to secrete a diverse repertoire of antibodies, with specificities for both foreign and auto-antigens. It is thought that the function of these cells may be to limit and control inflammation during the course of an ongoing immune response. A great deal of work is still required to tease out the lineage relationships among these various IL-10-secreting and other B-cell subpopulations. S U M M A R Y ■ ■ ■ ■ ■ ■ ■ The Clonal Selection Hypothesis states that each B cell bears a single antigen receptor. When antigen interacts with a B-cell receptor (BCR) specific for that antigen, it induces proliferation of that B cell to form a clone of cells bearing the identical specificity. Cells from this clone secrete antibodies having the same specificity as the antigen receptor. At the close of the immune response, there will be more cells bearing that specificity than existed prior to the response, and these so-called “memory B cells” will generate a faster and higher response upon secondary antigen exposure. Those cells bearing receptors with specificity for self antigens are eliminated from the B-cell repertoire during development. The B-cell response to some antigens requires help from T cells. These T-dependent responses are mounted by B-2 or follicular B cells. The response to some other antigens can occur in the absence of T-cell help. These T-independent responses are primarily mounted by B-1 or marginal zone (MZ) B cells. Mature B-2 B cells migrate to the lymphoid follicles under the influence of the chemokine CXCL13. Survival of mature B-2 B cells depends upon access to the survival factor BAFF. B cells may acquire antigens directly. Some antigens enter the lymph node by squeezing between the subcapsular sinus (SCS) macrophages that line the lymph node, whereas others enter via a leaky network of conduits that are sampled by the follicular B cells. Yet others are taken up first by the SCS macrophages or follicular dendritic cells and passed on to the B cells. When the BCR recognizes its cognate antigen, the receptors on the B-cell membrane briefly spread over the antigen surface, and then contract, resulting in B-cell receptor clustering. This represents the earliest phase of B-cell activation. Receptor clustering results in internalization of the receptorantigen complex followed by antigen presentation by the B cell to the T cell. The T-cell/B-cell interaction occurs primarily at the border between the T- and B-cell zones of the lymph node. T cells interact with their cognate B cells by binding to the processed antigen with their T-cell receptor (TCR), as well as by interactions between T cell CD28 with B cell CD80 and CD86, and between T cell CD40L and B cell CD40. These interactions facilitate the directional secretion of T-cell cytokines that are necessary for full B-cell activation. ■ ■ ■ ■ ■ ■ ■ ■ Following stimulation of primary B cells at the T-cell/ B-cell border within the lymph node, some B cells differentiate quickly into plasma cells that form primary foci and secrete an initial wave of IgM antibodies. This requires the up-regulation of the plasma cell transcription factors IRF4 and BLIMP-1. Other B cells from the antigen-stimulate clones migrate to the primary follicles and form germinal centers. Within germinal centers, B-cell differentiation continues, with the generation of somatically hypermutated receptors that are then subject to antigen selection, with the eventual generation of high-affinity antibodies. Somatic hypermutation (SHM) affects certain sequences called mutational hot spots, in the variable region genes of antibody molecules. Also within the germinal center, the constant region of the heavy chain of the antibody genes undergoes class switch recombination (CSR). This results in the formation of antibodies bearing mutated variable regions and constant regions other than ␮. CSR occurs between switch regions upstream of each heavy chain constant region gene (except for the ␦ constant region). Both somatic hypermutation and class switch recombination are mediated by the enzyme activation-induced cytidine deaminase (AID), followed by DNA repair. At the close of the B-cell response, long-term memory exists in two forms. Recirculating B memory cells must be reactivated by antigen in order to yield a higher, faster, and stronger response than the primary response. In addition, long-lived plasma cells residing in the bone marrow and other locations continually secrete antibodies and ensure that antibodies to commonly encountered antigens constantly circulate within the blood. T-independent responses generated by B-1 and marginal zone (MZ) B cells give rise to relatively low-affinity, primarily IgM antibodies. B-Cell Activation, Differentiation, and Memory Generation ■ TI-1 antigens interact with B cells via both the BCR and innate immune receptors, whereas TI-2 receptors are highly polymerized antigens that do not have an intrinsic, mitogenic activity. Both types of T-independent responses are enhanced by interactions with other cell types, including T cells, macrophages, and monocytes, and possibly, neutrophils. ■ ■ | CHAPTER 12 413 B-10 B cells release the interleukin IL-10 upon antigenic stimulation, and may serve to reduce inflammation during an ongoing immune response. B cells can be negatively signaled through CD22 and Fc␥RIIb—a receptor that recognizes the presence of IgGcontaining immune complexes in the blood. R E F E R E N C E S Ada, G. L., and P. Byrt. 1969. Specific inactivation of antigenreactive cells with 125I-labelled antigen. Nature 222:1291–1292. Allen, C. D., T. Okada, and J. G. Cyster. 2007. Germinal-center organization and cellular dynamics. Immunity 27:190–202. Allen, C. D., T. Okada, H. L. Tang, and J. G. Cyster. 2007. Imaging of germinal center selection events during affinity maturation. Science 315:528–531. Berek, C., G. M. Griffiths, and C. Milstein. 1985. Molecular events during maturation of the immune response to oxazolone. Nature 316:412–418. Burnet, F. M. 1957. A modification of Jerne’s theory of antibody production using the concept of clonal selection. The Australian Journal of Science 20:67–69. Catron, D. M., et al. 2004. Visualizing the first 50 hr of the primary immune response to a soluble antigen. Immunity 21:341–347. Dal Porto, J. M., et al. 2004. B cell antigen receptor signaling 101. Molecular Immunology 41:599–613. Dykstra, M., et al.2003. Location is everything: Lipid rafts and immune cell signaling. Annual Review of Immunology 21:457–481. Eisen, H. N., and G. W. Siskind. 1964. Variations in affinities of antibodies during the immune response. Biochemistry 3:996–1008. Flajnik, M. F., and M. Kasahara. 2010. Origin and evolution of the adaptive immune system: Genetic events and selective pressures. Nature Reviews Genetics 11:47–59. Harwood, N. E., and F. D. Batista. 2009. The antigen expressway: Follicular conduits carry antigen to B cells. Immunity 30:177–179. Hayakawa, K., R. R. Hardy, D. R. Parks, and L. A. Herzenberg. 1983. The “Ly-1 B” cell subpopulation in normal immunodefective, and autoimmune mice. Journal of Experimental Medicine 157:202–218. Jacob, J., and G. Kelsoe. 1992. In situ studies of the primary immune response to (4-hydroxy-3-nitrophenyl)acetyl. II. A common clonal origin for periarteriolar lymphoid sheathassociated foci and germinal centers. Journal of Experimental Medicine 176:679–687. Jacob, J., J. Przylepa, C. Miller, and G. Kelsoe. 1993. In situ studies of the primary immune response to (4-hydroxy-3nitrophenyl)acetyl. III. The kinetics of V region mutation and selection in germinal center B cells. Journal of Experimental Medicine 178:1293–1307. Kohler, G., and C. Milstein. 1975. Continuous cultures of fused cells secreting antibody of predefined specificity. Nature 256:495–497. Koshland, M. E. 1975. Structure and function of the J chain. Advances in Immunology 20:41–69. LeBien, T. W., and T. F. Tedder. 2008. B lymphocytes: How they develop and function. Blood 112:1570–1580. Martin, F., and J. F. Kearney. 2002. Marginal-zone B cells. Nature Reviews Immunology 2:323–335. Fleire, S. J., et al. 2006. B cell ligand discrimination through a spreading and contraction response. Science 312:738–741. Mitchell, G. F., and J. F. Miller. 1968. Cell to cell interaction in the immune response. II. The source of hemolysin-forming cells in irradiated mice given bone marrow and thymus or thoracic duct lymphocytes. Journal of Experimental Medicine 128:821–837. Germain, R. N., et al. 2008. Making friends in out-of-the-way places: How cells of the immune system get together and how they conduct their business as revealed by intravital imaging. Immunological Reviews 221:163–181. Muramatsu, M., et al. 2007. Discovery of activation-induced cytidine deaminase, the engraver of antibody memory. Advances in Immunology 94:1–36. Haas, K. M., J. C. Poe, D. A. Steeber, and T. F. Tedder, T.F. 2005. B-1a and B-1b cells exhibit distinct developmental requirements and have unique functional roles in innate and adaptive immunity to S. pneumoniae. Immunity 23:7–18. Hannum, L. G., A. M. Haberman, S. M. Anderson, and M. J. Shlomchik. 2000. Germinal center initiation, variable gene region hypermutation, and mutant B cell selection without detectable immune complexes on follicular dendritic cells. Journal of Experimental Medicine 192:931–942. Pape, K. A., D. M. Catron, A. A. Itano, and M. K. Jenkins. 2007. The humoral immune response is initiated in lymph nodes by B cells that acquire soluble antigen directly in the follicles. Immunity 26:491–502. Raff, M. C., M. Feldmann, and S. De Petris. 1973. Monospecificity of bone marrow-derived lymphocytes. Journal of Experimental Medicine 137:1024–1030. Schneider, P. 2005. The role of APRIL and BAFF in lymphocyte activation. Current Opinion in Immunology 17:282–289. 414 PA R T V | Adaptive Immunity: Effector Responses Shapiro-Shelef, M., and K. Calame. 2005. Regulation of plasma-cell development. Nature Reviews Immunology 5:230–242. Sixt, M., et al. 2005. The conduit system transports soluble antigens from the afferent lymph to resident dendritic cells in the T cell area of the lymph node. Immunity 22:19–29. Weigert, M. G., I. M. Cesari, S. J. Yonkovich, and M. Cohn. 1970. Variability in the lambda light chain sequences of mouse antibody. Nature 228:1045–1047. Weill, J. C., S. Weller, and C. A. Reynaud. 2009. Human marginal zone B cells. Annual Review of Immunology 27:267–285. Useful Websites http://bio-alive.com/seminars/immunology. htm This Bio Alive Web site is an excellent source of lec- loaded for personal use. The series contains lectures pertinent to this chapter, including seminars entitled Immunological Memory; Regulating B Cell Immunity; Moviemaking and Modeling (this one by Ronald Germain, one of the pioneers of the application of sophisticated imaging technology to the study of the immune system); Molecular Mechanisms of Leukocyte Migration; and Somatic Hypermutation. www.sciencedirect.com/science/ar ticle/pii/ S1074761304002389 Two movies from a paper by D. M. Catron et al. (2004, Visualizing the first 50 hr of the primary immune response to a soluble antigen, Immunity 21:341–347) that is discussed in Chapter 14 can be found at this Web site. The first portrays the movements of T cells and dendritic cells prior to antigen entry into the observed lymph node. The second shows the movements of T cells, B cells, and dendritic cells through a nearby lymph node, following injection of antigen into the footpad of a mouse. tures by accomplished immunologists that can be down- S T U D Y Q U E S T I O N S 1. Name one distinguishing feature of TI-1 and TI-2 antigens. 2. In the following flow cytometric dot plots, draw a circle where you expect to see the designated cell populations. (a) B2 (follicular) B cells (b) B-1 B cells 4. You are looking at lymph node sections of a mouse that has had its Lymphotoxin-␣ gene disrupted, and you are comparing it with sections from a mouse with an intact Lymphotoxin-␣ gene. Both mice were immunized approximately 10 days previously with a T-dependent antigen. What is the most striking feature that is missing in your knockout animal? CD5 CD5 5. You have immunized two mice with a T-dependent anti- IgM (c) B2 (follicular) B cells IgM (d) MZ B cells gen. One of them is a wild-type mouse, and the other belongs to the same strain as the first but is a knockout mouse that does not have the gene encoding activationinduced cytidine deaminase (AID). Upon primary immunization, both mice express similar titer (concentrations) of IgM antibodies. You allow the mice to rest for 6 weeks and then reimmunize. Do you expect the secondary antibodies to be similar or different between the two animals? Explain your answer. 6. Draw the biochemical reaction catalyzed by AID, and CD21 CD21 describe how subsequent DNA repair mechanisms can lead to somatic hypermutation (SHM) and to class switch recombination (CSR). 7. Describe one mechanism by which higher-affinity B cells may gain a selective survival advantage within the germinal center. 8. The presence of circulating antigen-antibody complexes IgM IgM 3. You have generated a knockout mouse that does not express the CD40 antigen. On stimulation with a T-dependent antigen, do you expect this mouse to be able to: a. secrete IgG antibodies specific for the antigen? b. secrete antigen-specific antibodies that have undergone somatic hypermutation? has been shown to result in the down-regulation of B-cell activation. State why this would be advantageous to the organism, and offer one mechanism by which the downregulation may occur. 9. Recent evidence suggests that not all regulatory cytokines secreted by lymphocytes derive from T cells. Explain. 10. Describe three mechanisms by which antigen can enter the lymph node and make contact with the B-cell receptor. 13 Effector Responses: Celland Antibody-Mediated Immunity I n previous chapters, we have focused on how the immune response is initiated. Here, we will finally describe how the targets of the immune system— pathogen, infected cells, and even tumor cells—are actually cleared from the body. We have already seen how the innate immune system initiates the response to pathogen and alerts the adaptive immune system to the presence and nature of that pathogen (Chapter 5). In Chapters 9 through 12, we described the development and activation of the antigenspecific cells of the adaptive immune system, B and T lymphocytes. You were also introduced to the differentiation and activity of helper T lymphocytes, a type of effector cell that regulates the activity and function of cytotoxic T cells, B cells, and other antigenpresenting cells. Here we focus on the effector cells and molecules of both the cell-mediated and the humoral (antibody-mediated) immune responses that directly rid the organism of pathogens and abnormal cells. These effector responses are arguably the most important manifestations of the immune response: they protect the host from infection and rid the host of pathogens that have breached defenses. The effector functions of cell-mediated and humoral branches of the immune system assume different, although overlapping, roles in clearing infection from a host. The effector molecules of the humoral branch are antibodies, the secreted version of the highly specific receptor on the surface of B cells. Antibodies secreted into extracellular spaces are exquisitely antigen specific and have several methods at their disposal to rid a body of pathogen. How an antibody contributes to clearing infection depends on its isotype, which determines whether it can recruit complement (recall Chapter 6). The isotype also determines which receptors an antibody can bind. Antibody-binding receptors, which bind to the constant regions of antibodies and are therefore called Fc receptors or FcRs, determine which cells an antibody can recruit to aid in its destructive mission, as well as the tissues to which it can gain entry. Two cytotoxic T lymphocytes bind to a tumor-specific antigen on the surface of a cancer cell, deliver the “kiss of death,” and induce apoptosis. [Steve Gschmeissner/Photo Researchers] ■ Antibody-Mediated Effector Functions ■ Cell-Mediated Effector Responses ■ Experimental Assessment of Cell-Mediated Cytotoxicity If antibodies were the only agents of immunity, intracellular pathogens, which occupy spaces that antibody cannot access, would likely escape the immune system. Fortunately there is another effector branch of our immune system, cell-mediated immunity, which detects and kills cells that harbor intracellular pathogens. Cellmediated immunity consists of both helper T cells (CD4⫹ TH) and several types of cytotoxic cells. As you have seen in Chapter 11, TH cells exert their effector functions indirectly, by contributing to the activation of antigen-presenting cells, B cells, and cytotoxic T cells via receptor-ligand interactions and soluble cytokines and chemokines. On the other hand, cytotoxic cells exert their effector functions directly, by attacking infected cells and, in some cases, the pathogens themselves. Effector cytotoxic cells arise from both the adaptive and innate immune systems and, therefore, include both antigen-specific and -nonspecific cells. Antigennonspecific (innate immune) cells that contribute to the 415 416 PA R T V | Adaptive Immunity: Effector Responses clearance of infected cells include NK cells and nonlymphoid cell types such as macrophages, neutrophils, and eosinophils. Antigen-specific cytotoxic cells include CD8⫹ T lymphocytes (CTLs or TC cells), as well as the CD4⫹ NKT cell subpopulation, which, although derived from the T-cell lineage, displays some useful features of innate immune cell types, too. Populations of cytotoxic CD4⫹ TH cells have also been described and may contribute to delayed-type hypersensitivity. The discussion of DTH reactions and the role of CD4⫹ T cells in their orchestration appears in Chapter 15. The humoral and cell-mediated immune systems also cooperate effectively. Cells such as macrophages, NK cells, neutrophils, and eosinophils all express Fc receptors, which induce phagocytosis of antibody-antigen complexes as well as direct killing of target cells via a process known as antibody-dependent cell-mediated cytotoxicity. It is also important to note that the cell-mediated immune system plays a role in recognizing and eliminating not only infected cells, but tumor cells, which often have undergone genetic modifications that lead to surface expression of antigens not typical of normal cells. In this chapter we will first focus on antibodymediated activities, describing not only their ability to fix complement but also the multiple functions they acquire by interacting with cells expressing Fc receptors. We will then describe cytotoxic effector mechanisms mediated by cells of the innate and adaptive immune system. We will close with a discussion of experimental assays of cytotoxicity. A Clinical Focus feature and a Classic Experiment feature are also included in this chapter. The first describes how knowledge of antibody structure and function has led to new and successful disease therapies. The second introduces recent data showing that lymphocytes are not the only cell type that can develop a memory; natural killer cells, members of the innate immune system, also appear to have this capacity. Antibody-Mediated Effector Functions Antibodies generated by activated B cells (see Chapter 12) protect the body against the pathogen in several ways: they can neutralize pathogen by binding to and blocking receptors that the pathogen uses to gain entry into a cell; they can opsonize pathogen by binding and recruiting phagocytic cells; and they can fix complement by binding to pathogen and initiating the complement cascade, which punctures cell membranes (see Chapter 6). In addition, they can cooperate with the cellular branch of the immune system by binding and recruiting the activities of cytotoxic cells, specifically natural killer (NK) cells, in a process called antibody-dependent cell-mediated cytotoxicity (ADCC) (Figure 13-1 and Table 13-1). The ability of antibodies to mediate these responses is dependent not only on their antigen specificity, but also on their isotype (see Chapter 3 and Table 13-2), which determines whether an antibody will “fix” complement, to which Fc receptors they will bind, and therefore, which cellular effects they will have. Collectively, these mechanisms result in the destruction of pathogen, either directly or indirectly, by stimulating phagocytosis and digestion of pathogen-antibody complexes, directly destroying pathogenic cell membranes, or inducing suicide of the infected cell. Antibodies Mediate the Clearance and Destruction of Pathogen in a Variety of Ways Antibodies are versatile effector molecules that play a direct role in resolving infection by (1) blocking pathogen entry into cells (neutralization) and (2) recruiting cytotoxic molecules and cytotoxic cells to kill pathogen (via complement fixation, opsonization, and ADCC). Neutralization Most viruses and some bacteria gain entry into a cell by binding specifically to one or more cell-surface proteins, stimulating endocytosis. For example, the human immunodeficiency virus (HIV) expresses a coat protein, gp120, that specifically binds the CD4 coreceptor. The influenza virus expresses a protein, hemagglutinin, that binds to sugarmodified proteins on epithelial cells. Antibodies that bind such proteins are particularly potent effector molecules because they can prevent a pathogen from ever initiating an infection. They are referred to as neutralizing antibodies. Pathogens disarmed by neutralizing antibodies are typically phagocytosed by macrophages. Neutralizing antibodies can also block entry of toxins into cells. For example, tetanus toxoid, a product of the Clostridium tetani bacteria, is a neurotoxin that can result in uncontrolled muscle contraction and death. The tetanus vaccine contains an inactive version of this toxin that stimulates B-cell production of anti-tetanus toxoid antibodies. These antibodies bind and potently inhibit the entry of the toxin into nerve cells. Neutralizing antibodies have also been raised against snake venom toxins and are effective therapies for some snakebites. Physiologically, neutralization is the most effective mode of protection against infection. However, because microbes proliferate rapidly, they can generate genetic Effector Responses: Cell- and Antibody-Mediated Immunity | CHAPTER 13 417 Variable regions (Fab) Constant regions (Fc) C1q C1s C1r NK cell FcR (1) Virus and toxin neutralization Tumor cell Macrophage MAC (2) Opsonization Prevents pathogen-host binding Phagocytosis Infected or tumor cell (3) Complement fixation and formation of the membrane attack complex (4) Antibody-Dependent Cell-mediated Cytotoxicity (ADCC) Phagocytosis or lysis NK-induced apoptosis functions. Antibody binding can enhance the clearance of pathogens by (1) binding to proteins that pathogens use to infect cells (neutralization), (2) interacting with Fc receptors on the surface of phagocytic cells and inducing internalization and degradation (opso- nization), (3) recruiting complement proteins that can directly kill pathogen or enhance its phagocytosis, and (4) binding to FcRs on cytotoxic cells (e.g., NK cells) and directing their activity to infected cells, tumor cells, and/or pathogens (antibody-dependent cell-mediated cytotoxicity [ADCC]). variants able to evade neutralizing antibodies. For example, influenza viral variants expressing modified hemagglutinins that do not bind circulating antibodies have a distinct reproductive advantage and can rapidly take over an infection. HIV is one of the most adept viruses at evading anti- body responses. In theory, gp120 is an ideal target for neutralizing antibodies; however, few individuals generate such antibodies. This is in part because the region of gp120 that binds CD4 is concealed from the body until just before viral entry. FIGURE 13-1 Four broad categories of antibody effector TABLE 13-1 Ways antibodies can protect the host against invasion by pathogens Mode of antibody-mediated protection How it works Neutralization Antibody binds to sites on the pathogen or a toxin that interact with host proteins, masking them, and inhibiting entry of that pathogen or toxin into the host. Antibody-pathogen complexes are then eliminated, often after phagocytosis. Opsonization Antibody binds to pathogen and is then bound by Fc receptors on phagocytic cells. The antibodyantigen binding to FcRs induces internalization and destruction by the phagocytic cell. Complement fixation The antibody-antigen complex becomes bound by complement components in serum and is either phagocytosed via cells expressing C3 receptors or lysed as a result of pore formation by the complement components C7, C8, and C9. Antibody-dependent cell-mediated cytotoxicity Antibody-antigen complexes are bound by Fc receptors on NK cells and granulocytes, thus directing the cytotoxicity of these cells toward the antigen targeted by the antibody (e.g., viral proteins on the surface of an infected cell), and inducing apoptosis of the target cell. 418 PA R T V TABLE 13-2 | Adaptive Immunity: Effector Responses Properties and biological activities* of classes and subclasses of human serum immunoglobulins IgG1 IgG2 IgG3 IgG4† IgA1 IgA2 IgM§ IgE IgD Molecular weight‡ 150,000 150,000 150,000 150,000 150,000– 600,000 150,000– 600,000 900,000 190,000 150,000 Heavy-chain component ␥1 ␥2 ␥3 ␥4 ␣1 ␣2 ␮ ␧ ␦ Normal serum level (mg/ml) 9 3 1 0.5 3.0 0.5 1.5 0.0003 0.03 In vivo serum half-life (days) 23 23 8 23 6 6 5 2.5 3 Activates classical complement pathway ⫹ ⫹/⫺ ⫹⫹ ⫺ ⫺ ⫺ ⫹⫹ ⫺ ⫺ Crosses placenta ⫹ ⫹Ⲑ⫺ ⫹ ⫹ ⫺ ⫺ ⫺ ⫺ ⫺ Present on membrane of mature B cells ⫺ ⫺ ⫺ ⫺ ⫺ ⫺ ⫹ ⫺ ⫹ Binds to Fc receptors of phagocytes ⫹⫹ ⫹Ⲑ⫺ ⫹⫹ ⫹ ⫺ ⫺ ? ⫺ ⫺ Mucosal transport ⫺ ⫺ ⫺ ⫺ ⫹⫹ ⫹⫹ ⫹ ⫺ ⫺ Induces mast cell degranulation ⫺ ⫺ ⫺ ⫺ ⫺ ⫺ ⫺ ⫹ ⫺ * Activity levels are indicated as follows: ⫹⫹ ⫽ high; ⫹⫽ moderate; ⫹/⫺ ⫽ minimal; ⫺⫽ none; ? ⫽ questionable. Note that mice do not make IgG4 but generate two different versions of IgG2 (IgG2a and IgG2b). ‡ IgG, IgE, and IgD always exist as monomers; IgA can exist as a monomer, dimer, trimer, or tetramer. Membrane-bound IgM is a monomer, but secreted IgM in serum is a pentamer. § IgM is the first isotype produced by the neonate and during a primary immune response. † Opsonization From the Greek word for “to make tasty,” opsonization refers to the ability of antibodies to promote and/or enhance the engulfment of antigens by phagocytes. Opsonizing antibodies coat antigen and interact with Fc receptors expressed by phagocytic cells. Thus, phagocytic cells of the innate immune system (e.g., macrophages and neutrophils) have two sets of receptors capable of binding pathogens: they can bind them directly using innate immune receptors (pattern recognition receptors such as Toll-like receptors), or they can bind antigens bound in immune complexes with antibodies, via Fc receptors. Innate immune receptors are important in activating antigen-presenting cells and shaping the signals that they will convey to lymphocytes (see Chapter 5 and Chapter 11). Fc receptors also generate signals and regulate the effector responses of innate immune cells (see below). In the case of opsonization, binding of pathogen (antigen)-antibody complexes to an Fc receptor on phagocytes will induce internal- ization of the complex and internal digestion of the pathogen in lysosomes. Complement Fixation As we saw in Chapter 6, antigen-antibody complexes can also induce a complement cascade. Specifically, when antibodies that associate with complement bind to the surface of bacteria and some (enveloped) viruses, they can initiate a cascade of reactions that results in the generation of the membrane attack complex (MAC), perforating pathogen membranes and killing the microbe. This capacity to stimulate complement-mediated membrane damage is referred to as fixing complement and is a property of specific antibody isotypes, including some IgGs and IgM. Activation of the early part of the complement cascade can also protect the host by opsonizing pathogens in an antibodyindependent manner. This complement function does not require the later components (C4–C9) of the complement cascade. Instead, the complement protein fragment C3b, which is produced early on in the complement reaction, binds to the Effector Responses: Cell- and Antibody-Mediated Immunity | CHAPTER 13 419 surface of the pathogen and is recognized by the many different blood cells that have C3b receptors. Binding of C3b on the surface of a pathogen by a macrophage results in phagocytosis and destruction of that pathogen by mechanisms similar to those described above for antibody-mediated phagocytosis. Interestingly, binding of C3b-bound pathogen by red blood cells results in the delivery of the pathogens to the liver or spleen, where the pathogen is removed from the red blood cell, without affecting the viability of the red blood cell, followed by phagocytosis of the pathogen by a macrophage. are often pentavalent (bound together in groups of five) they can be very effective in binding pathogen. Interestingly, most circulating IgM antibodies come not from conventional B cells but from B-1 cells, which constitutively produce IgM antibodies that appear to protect us from common pathogens, particularly those that can come from the gut and other mucosal areas. IgM antibodies are very good at fixing complement and efficiently induce lysis of pathogens to which they bind. They are also good at forming dense antibody-pathogen complexes that are efficiently engulfed by macrophages. Antibody-Dependent Cell-Mediated Cytotoxicity Antibodies can recruit the activities of multiple cytotoxic cells, including NK cells (lymphoid members of the innate immune system) and granulocytes (myeloid members of the innate immune system). Unlike cytotoxic T lymphocytes, these cells do not express antigen-specific T-cell receptors. However, they do express Fc receptors (specifically, Fc␥RIII) and can use these to “arm” themselves with antibodies that allow them to “adopt” an antigen specificity. Antibodies bound to NK cells then direct the cytotoxic activity to a specific cellular target—for example, an infected cell that expresses viral proteins on its surface or a tumor cell that expresses tumorspecific proteins on its surface (see Figure 13-1). This phenomenon, referred to as ADCC, is under intense investigation because of its therapeutic potential. Monoclonal antibodies (mAbs) developed as cancer treatments are thought to act in part by taking advantage of ADCC mechanisms and directing NK cytotoxicity toward tumor cell targets (see Clinical Focus Box 13-1). As with complement fixation, the ability to mediate ADCC is antibody-isotype dependent, and murine IgG2a and human IgG1 are most potent in this capacity (see below). Effector Functions of IgG Antibodies IgG antibodies are the most common antibody isotype in the serum. They are also the most diverse, and include several subclasses (IgG1, IgG2, IgG3, and IgG4 in humans; IgG1, IgG2a, IgG2b, and IgG3 in mice), each of which has distinct effector capabilities. All of the IgG variants bind to Fc receptors and can enhance phagocytosis by macrophages (opsonization). Human IgG1 and IgG3 are particularly good at fixing complement. Mouse IgG2a and human IgG1 are particularly good at mediating ADCC by NK cells. Specific IgG isotypes are also associated with disease pathologies. For example, in a mouse lupus model, IgG2a and IgG2b isotypes dominate. In human lupus, IgG1 and IgG3 are involved in anti-DNA reactions, and IgG2 and IgG3 are associated with kidney dysfunction. Investigators are making excellent use of our growing knowledge of IgG effector functions and are tailoring the isotypes of mAbs used in therapies to enhance their effect. For example, IgG1 antibodies are most commonly used in tumor therapy because they can both fix complement and mediate ADCC by NK cells. However, recent data suggest that IgG2 is a potent mediator of ADCC by myeloid cells (including neutrophils) and may offer some therapeutic advantages. Antibody Isotypes Mediate Different Effector Functions As you learned in Chapters 3 and 12, activated B cells develop into plasma cells, which are remarkably productive antibodyproducing cellular factories. Antibody specificity is defined by their Fab regions, and their isotype or class is defined by their Fc region (see Figure 3-21). Depending on the type of T cell help they received during activation, as well as the cytokines to which they were exposed, B cells will secrete one of four major classes of antibody: IgM, IgG, IgA, or IgE. Although immature B cells express IgD, little if any is secreted and its function remains a bit of a mystery. The structure of each of these antibodies is reviewed in detail in Chapter 3. Each plays a distinct role in thwarting and clearing infection, as described below and depicted in Figure 13-2. Effector Functions of IgM Antibodies IgM antibodies are the first class of antibodies to be produced during a primary immune response. They tend to be lowaffinity antibodies because the B cells that produce them have not gone through affinity maturation. However, because they Effector Functions of IgA Antibodies Although IgA antibodies are also found in circulation, they are the major isotype found in secretions, including mucus in the gut, milk from mammary glands, tears, and saliva. In these secretions, IgA can neutralize both toxins and pathogens, continually interacting with the resident (commensal) bacteria that colonize our mucosal surfaces and preventing them from entering the bloodstream. Because IgA cannot fix complement, these interactions do not induce inflammation, which is advantageous given that IgA acts continually on antigens and pathogens that typically pose no threat. IgA’s half-life in secretions is relatively long because the amino acid sequence of the Fc region is resistant to many of the proteases that are present. Interestingly, several microbes, including those that cause gonorrhea and strep infections, produce proteases that do degrade IgA, allowing them to evade this form of immune protection. IgA does exist in monomeric form (particularly in circulation) but also forms dimers and polymers in mucosal tissues. The dimeric and polymeric forms bind to receptors 420 PA R T V | Adaptive Immunity: Effector Responses CLINICAL FOCUS Monoclonal Antibodies in the Treatment of Cancer Cancer is a disease in which a single cell multiplies rapidly and uncontrollably to form a tumor. As the tumor cells replicate, the DNA repair functions that normally ensure faithful replication of DNA sequence during cell division may begin to fail, and new mutations begin to accumulate, leading to a worsening prognosis. Cancer remains one of the worst killers in the developed world, and despite decades of intensive and creative work by research scientists and clinicians, a diagnosis of cancer is still followed all too frequently by a shortened life span, preceded by a course of weakening and toxic treatment. Attempts to develop effective anticancer drug strategies have had to address two main problems. First, because many cancer cells are derived from self and do not “look” too different from healthy cells, any drug that binds to the cancer cell is likely also to bind and enter the patient’s other, healthy cells. The trick is to find a molecule on the surface of a cancer cell that is either not expressed in normal cells, or is expressed at much lower concentrations in normal cells compared to cancer cells, and then to design a drug that will preferentially bind to the malignant cell. Second, because cancer cells and normal dividing cells use the same metabolic strategies to generate ATP, any untargeted drug that affects the metabolism or growth of a cancer cell will also affect normal dividing cells. Most of the first generation of anticancer drugs, however, are focused on inhibiting cancer cell division. Drugs such as daunorubicin, cis-platin, and others enter all cells of the body and interfere with the process of cell division. As mentioned above, such drugs will also target cells that divide normally as part of their dayto-day function, and the patient’s wellbeing will be compromised not only by the tumor they carry, but also by the treatment they endure. Cells of the alimentary tract, skin, hair follicles, and immune system all divide more frequently than others such as the liver, kidney, and brain and tend to be more sensitive to these drugs. This is why conventional chemotherapy leads to hair loss, digestive distress, and impaired immune function. The development of monoclonal antibodies, with their extraordinary capacity to target a single 6- to 8-amino-acid length of a protein was seized upon by scientists as a potential opportunity to specifically target tumor cells, while leaving the neighboring healthy cells undamaged. Immunotherapeutics, the production and refinement of mAbs that will specifically target cancer cells for death and healthy neighbors, is a burgeoning field. Now part of the therapeutic arsenal, mAbs have met with some success. Some examples of current reagents and their mechanisms of action are described below. How Do Antibodies Kill Cancer Cells? • Competitive binding to cell surface receptors and preventing their activation. Antibody binding to molecules on the surface of a cell can block interactions that may be necessary to allow growth of that cell. For example, many tumor cells become dependent for their growth on binding of specific growth factors such as Epidermal growth factor (EGF). Several of the mAbs currently in clinical use, such as erbitux, bind to the EGF receptor (EGFR) and interfere with its function. Herceptin is an example of an extremely successful drug that targets the Her-2 EGFR on advanced breast cancer cells. All EGFRs must dimerize on binding to EGF so that they can send their growth signal to the nucleus. Herceptin interferes with this dimerization reaction and, so, interrupts the EGFR signal. • Interference with the generation of new blood vessels. Another group of drugs, including bevacizumab (Avastin), targets activity of vascular endothelial growth factor (VEGF). VEGF interacts with VEGF receptors (VEGFR1 and VEGFR2) located on blood vessels. Binding stimulates the formation of new blood vessels that can be used to supply the nascent tumor cells with nutrients. Avastin binds to VEGF and prevents it from interacting with its receptors, thereby inhibiting the development of new blood vessels and compromising tumor growth. • Direct binding to receptors with induction of apoptosis. Some mAbs, such as rituximab (Rituxan), may kill their target cells by binding to the surface of the cell and mimicking the binding of natural ligands. Binding of rituximab to follicular B lymphoma cells results in interference with cell cycle regulation and the induction of apoptosis. • Induction of ADCC. Binding of certain isotypes of antibodies to the surface of a cell results in the induction of antibody-dependent cell-mediated cytotoxicity (ADCC). If a tumor cell becomes coated with antibodies of human IgG1 or IgG3 subclasses, Fc␥RIII molecules on the surface of natural killer (NK) cells and other immune cells will bind to the Fc regions of these antibodies. Once bound to the NK cell, the tumor cell becomes the target of secreted perforin and granzymes, which induce apoptosis in the affected cell. In addition, macrophages will recognize the Fc portion of antibody clusters of the IgG1 isotype that are bound to a cell surface, and be induced to phagocytose the antibody-cell complex and/or release toxic metabolites. • Induction of complement fixation. IgG3 and IgG1 antibody binding to the surface of a cell can also result in the attachment (fixation) of a set of serum proteins known as the complement cascade to the cell surface. By a complex set of reactions detailed in Chapter 6, this can result in either the dissolution of the cell following disruption of the integrity of its membrane, or enhanced phagocytosis of that cell by macrophages. • Delivery of toxins. Very early on in the development of mAbs as clinical tools, Effector Responses: Cell- and Antibody-Mediated Immunity | CHAPTER 13 421 BOX 13-1 scientists began modifying them with toxins such as radioactive isotopes and cytotoxic reagents. The isotopes 90Y and 131 I are both capable of delivering shortrange cytotoxic doses of radiation to anything bound by the conjugated antibody and the immuno-conjugates Zevalin and Bexxar were approved for use in humans for the treatment of relapsed or refractory non-Hodgkin’s lymphoma in 2002 and 2003, respectively. Other toxins that have been studied in the lab and/or used in the clinic include modified bacterial toxins such as Pseudomonas exotoxin A, which has been conjugated to an mAb directed toward the CD22 molecule and has proven promising in the treatment of chemoresistant hairy cell leukemia. Ricin, another bacterial toxin, has also TABLE 1 been successfully conjugated with antibodies directed toward CD22 and CD19, and these molecules are currently being developed for the treatment of pediatric and acute lymphoblastic leukemia. Currently, scientists are working to improve the biological efficacy of the antibodies employed in cancer treatment by subjecting proven antibodies to mutation in vitro followed by selection using phage display technology to improve the affinity and selectivity of binding to tumor cells. Others are combining multiple antibodies with different specificity to the same cell in a single treatment, or combining mAbs with more conventional chemotherapeutic approaches. The Human Genome Project has provided a wealth of information regarding the number and structure of cell-surface signaling molecules, and it is also being exploited in the design of novel drugs and toxins, including antibodies and antibody fragments. For further reading on this topic, see the references below. Also keep an eye on the news. Immunotherapeutics is one of the fastest-moving and most exciting areas of research at the confluence of the laboratory and the clinic. New ideas arise every day, and new reagents are likely to follow. Rivera, F., et al. 2008. Current situation of panitumumab, matauzumab, nimotuzumab and zalutumumab. Acta Oncologia 47:9–19. Strome, S. E., et al. 2007. A mechanistic perspective of monoclonal antibodies in cancer beyond target-related effects. The Oncologist 12:1084–1095. Some monoclonal antibodies in clinical use mAb product (trade name) Nature of antibody Target (antibody specificity) Modification of antibody Rituximab, (Rituxan) Chimeric CD20 (mouse B-cell antigen) None Relapsed or refractory non-Hodgkin’s lymphoma Trastuzumab (Herceptin) Humanized Human epidermal growth factor receptor 2 (HER-2) None HER-2 receptor positive advanced breast cancers Alemtuzumab (Campath) Humanized CD52 (an antigen on many types of leukocytes) None Chronic lymphocytic leukemia Bevacizumab (Avastin) Humanized Vascular endothelial growth factor (VEGF) None Colorectal cancer Cetuximab (Erbitux) Chimeric EGFR None Colorectal cancer Panitumumab (Vectibix) Human EGFR None Colorectal cancer Ibritumomab Mouse CD20 None Relapsed or refractory non-Hodgkin’s lymphoma Ibritumomab tiuxetan (Zevalin) Mouse CD20 90 Relapsed or refractory non-Hodgkin’s lymphoma Tositumomab (Bexxar) Mouse CD20 131 Gemtuzumab Humanized ozogamicin (Mylotarg) CD33 (glycoprotein antigens on myeloid progenitor cells and monocytes) Attached to an anti-tumor agent that cleaves double-stranded DNA at specific sequences Acute myelogenous leukemia Epratuzumab CD22 (glycoprotein antigens on mature and neoplastic B cells) None Relapsed or refractory non-Hodgkin’s lymphoma Humanized Y I Treatment Relapsed or refractory non-Hodgkin’s lymphoma PA R T V 422 | Adaptive Immunity: Effector Responses FcR Eosinophil Helminth (a) Degranulation (FcεR and IgE) (b) Opsonization of bacteria and phagocytosis (FcγR and FcαR and IgG or IgA) Digestion in lysosome Phagocyte Blood (c) Maintaining serum levels of antibodies (FcRn and IgG) Plasma cell Dissociation at physiological pH Acidic endosome Endothelial cell (d) ADCC (FcγR and IgG) NK cell Tumor cell (e) Transcytosis into secretions (PolyIgR and IgA) Lumen (GI tract) Tumor Ag IgA Poly-Ig receptor Epithelial cell Blood Secreted IgA Secretory component FIGURE 13-2 Functions of Fc receptors. Fc receptors (FcRs) come in a variety of types and are expressed by many different cell types. Some of their major functions are illustrated here. (a) When bound by IgE-pathogen (e.g., worm) complexes, Fc␧Rs expressed by granulocytes can induce the release of histamine and proteases (degranulation). (b) When bound by antibody-pathogen complexes, Fc␣R and Fc␥R can induce macrophage activation and phagocytosis. (c) The neonatal FcR (FcRn) binds antibody that has been nonspecifically engulfed by endothelial cells and returns the antibody, intact, to the blood. (d) When bound by IgG antibodies coating infected or tumor cells, Fc␥Rs activate the cytolytic activity of natural (NK) cells. (e) PolyIg receptors (PolyIgR) expressed by the inner (basolateral) surface of epithelial cells (facing the blood) will bind dimers and multimers of IgA and IgM antibodies and transfer them through the cell to their apical (outer) surface and into the lumen of an organ (e.g., the GI tract). This is a process referred to as transcytosis and is responsible for the accumulation of antibodies in bodily secretions. Effector Responses: Cell- and Antibody-Mediated Immunity | CHAPTER 13 423 lysis, many of the effector functions of the antibodies described above depend on their ability to interact with receptors on cells in the innate and adaptive immune systems as well as cells in various epithelial and endothelial tissues. These antibody receptors were originally discovered over 40 years ago by early antibody biochemists, who referred to them as Fc receptors because they bound specifically to the Fc portion of the antibody. FcRs allow nonspecific immune cells to take advantage of the exquisite specificity of antibodies to focus their cellular functions on specific antigens and pathogens. They also provide a physical bridge between the humoral and cellmediated immune systems. When bound by antibody alone, FcRs do not trigger signals; however, when cross-linked by multiple antibody-antigen complexes, FcRs initiate effector responses (see Figure 13-2). The diversity of FcRs, most of which are members of the immunoglobulin superfamily (Figure 13-3), is now fully appreciated. Each differs by (1) the antibody class (isotype) to which they bind (e.g., Fc␥Rs bind to several IgG isotypes, Fc␧Rs bind to the IgE isotype, and Fc␣Rs bind to IgA isotypes), (2) the cells that express them (macrophages, granulocytes, NK cells, T and B cells), and (3) their signaling properties. Together these properties define a distinct effector function that an antibody can inspire (Table 13-3). on epithelial cells, an event that triggers endocytosis and transport of the molecules from the basolateral (inner) to the apical (outer) sides of the epithelial cell and into the tissue lumen (see Figure 13-2). Two subclasses of IgA—IgA1 and IgA2—are found in humans. IgA1 is more prevalent in the serum (and typically monomeric), and IgA2 is more prevalent in secretions. Both isotypes can mediate ADCC by binding FcRs on NK T cells, and both can trigger degranulation of granulocytes. Effector Functions of IgE Antibodies IgE antibodies are best known for their role in allergy and asthma. However, research suggests that they play an important role in protection against parasitic helminths (worms) and protozoa. They are made in very small quantities but have a very potent effect, inducing degranulation of eosinophils and basophils, and release of molecules such as histamine that do permanent damage to a large pathogen. Fc Receptors Mediate Many Effector Functions of Antibodies Although several antibody isotypes have a direct effect on the viability of a pathogen by initiating complement-mediated Poly IgR S S S S S S Fcγ RI FcRN S S S S S S S S S S S S β2m S S Fcγ RIIB Fcγ RIII FcαR FcεRI S S S S S S S S S S S S S S S S S S S S ITIM ITAM γ γ ITAM CD32 β γ γ CD16 CD64 ITAM γ γ ITAM β γ γ CD89 Inhibitory FIGURE 13-3 The structure of a number of human Fc receptors. The antibody-binding polypeptides are shown in blue and, where present, accessory signal-transducing polypeptides are shown in green. Most FcRs are activating receptors and are associated with signaling proteins that contain the Immuno-receptor Tyrosine Activation Motif (ITAM) or include ITAM(s) in their own intracellular regions. Fc␥RIIB is an inhibiting FcR and includes an Immunoreceptor Tyrosine Inhibition Motif (ITIM) in its intracellular region. All FcRs shown in this figure are members of the immunoglobulin supergene family; their extracellular regions include two or more immunoglobulin folds. These molecules are expressed on the surface of various cell types as cell-surface antigens and, as indicated in the figure, some have been assigned cluster of differentiation (CD) designations (for clusters of differentiation, see Appendix 1). [Adapted from M. Daeron, 1999, Fc Receptors, in The Antibodies, M. Zanetti and J. D. Capra, eds., vol. 5, p. 53, Amsterdam: Harwood Academic Publishers.] 424 PA R T V TABLE 13-3 | Adaptive Immunity: Effector Responses Expression and function of FcRs FcR Isotypes they bind Cells that express them Function Fc␥RI (CD64) IgG2a in mice, IgG1 and IgG3 in humans High-affinity receptor Dendritic cells, monocytes, macrophages, granulocytes, B lymphocytes Phagocytosis Cell activation Fc␥RII (CD32) IgG Dendritic cells, monocytes, macrophages, granulocytes, B lymphocytes, some immature lymphocytes Inhibitory receptor Traps antigen-antibody complexes in germinal center Abrogates B-cell activation Fc␥RIII (CD16) Humans generate two versions: Fc␥RIIIA (CD16a) and Fc␥RIIIB (CD16b) IgG1, IgG2a, and IgG2b in mouse; IgG1 in human Low -affinity receptor Only FcR that binds mouse IgG1 Dendritic cells, monocytes, macrophages, granulocytes, B lymphocytes Only FcR expressed by NK cells ADCC Cell activation Fc␥RIV (in mouse, with some similarity to human Fc␥RIIIA and/or human Fc␧RI) IgG2a and IgG2b in mice; IgG1 in Monocytes, macrophages, humans granulocytes Intermediate affinity receptor, Not on lymphocytes although exhibits higher affinity for human IgG1 than FcgRIIA. ADCC Cell activation Fc␧RI IgE Eosinophils, basophils, mast cells Degranulation of granulocytes, including eosinophils, basophils, mast cells Fc␧RII (CD23) IgE (low affinity) B lymphocytes Regulation of B-cell production of IgE Transport of IgE-antigen complexes to B-cell follicles Fc␣RI (CD89) IgA Dendritic cells, monocytes, macrophages, granulocytes, some liver cells Phagocytosis Cell activation ADCC pIgR IgA and IgM Multiple epithelial cells Transport of antibody from blood to the lumens of GI, respiratory, and reproductive tracts (transcytosis) FcRn (neonatal FcR) IgG Epithelial cells (including intestinal epithelium) Endothelial cells of mature animals Transport of antibodies from milk to blood (transcytosis) Transport of antibody-pathogen complexes from gut to mucosal immune tissue Phagocytosis Maintenance of levels of serum IgG and albumin For example, antibody-antigen complexes binding to Fc␧Rs on eosinophils and other granulocytes trigger the release of histamine and proteases from granules (degranulation). Fc␥Rs and Fc␣Rs on macrophages trigger a signaling cascade that induces the internalization of the pathogen into a phagocyte, which can destroy it via a variety of mechanisms, including oxidative damage, enzyme digestion, and membrane-disrupting effects of antibacterial peptides. “Neonatal” FcR (FcRn) molecules on cells that line the bloodstream help maintain antibody levels in serum by protecting them from degradation. Other Fc␥Rs expressed by cytotoxic cells, including NK cells