Jump to content

Electrochemical gradient: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
→‎Ion gradients: use original SVG
Copyedit, linking, added cleanup tags and templates.
(18 intermediate revisions by 17 users not shown)
Line 1: Line 1:
{{Short description|Gradient of electrochemical potential, usually for an ion that can move across a membrane}}
[[File:Membrane potential ions en.svg|thumb|Diagram of ion concentrations and charge across a semi-permeable cellular membrane.]]
[[File:Membrane potential ions en.svg|thumb|Diagram of ion concentrations and charge across a semi-permeable cellular membrane.]]


An '''electrochemical gradient''' is a gradient of [[electrochemical potential]], usually for an [[ion]] that can move across a [[membrane]]. The gradient consists of two parts:
An '''electrochemical gradient''' is a [[gradient]] of [[electrochemical potential]], usually for an [[ion]] that can move across a [[membrane]]. The gradient consists of two parts, the chemical gradient, or difference in [[Concentration|solute concentration]] across a membrane, and the electrical gradient, or difference in [[Electric charge|charge]] across a membrane. When there are unequal concentrations of an ion across a permeable membrane, the ion will move across the membrane from the area of higher concentration to the area of lower concentration through [[Molecular diffusion|simple diffusion]]. Ions also carry an electric charge that forms an [[electric potential]] across a membrane. If there is an unequal distribution of charges across the membrane, then the difference in electric potential generates a force that drives ion diffusion until the charges are balanced on both sides of the membrane.<ref name=":5">{{Cite book|title=Lehninger Principles of Biochemistry|last1=Nelson|first1=David|last2=Cox|first2=Michael|publisher=W.H. Freeman|year=2013|isbn=978-1-4292-3414-6|location=New York|pages=403}}</ref>


* The chemical gradient, or difference in [[Concentration|solute concentration]] across a membrane.
== Definition ==
* The electrical gradient, or difference in [[Electric charge|charge]] across a membrane.
The electrochemical gradient is the gradient of the electrochemical potential:
:<math>\nabla \overline{\mu}_i = \nabla \mu_i(\vec{r}) + \nabla z_i\mathrm{F}\varphi (\vec{r}) </math>, with


When there are unequal concentrations of an ion across a permeable membrane, the ion will move across the membrane from the area of higher concentration to the area of lower concentration through [[Molecular diffusion|simple diffusion]]. Ions also carry an electric charge that forms an [[electric potential]] across a membrane. If there is an unequal distribution of charges across the membrane, then the difference in electric potential generates a force that drives ion diffusion until the charges are balanced on both sides of the membrane.
* <math>\mu_i</math> the chemical potential of the ion species <math>i</math>

* <math>z_i</math> the valency of the ion species <math>i</math>
Electrochemical gradients are essential to the operation of [[Electric battery|batteries]] and other [[electrochemical cell]]s, [[photosynthesis]] and [[cellular respiration]], and certain other biological processes.
* F, [[Faraday constant]]
* <math>\varphi</math> the local [[electric potential]]


==Overview==
==Overview==
Electrochemical energy is one of the many interchangeable forms of [[potential energy]] through which energy may be [[conservation of energy|conserved]]. It appears in [[electroanalytical chemistry]] and has industrial applications such as batteries and fuel cells. In biology, electrochemical gradients allow cells to control the direction ions move across membranes. In [[mitochondria]] and [[chloroplast]]s, [[proton-motive force|proton gradients]] generate a '''chemiosmotic potential''' used to synthesize [[Adenosine triphosphate|ATP]],<ref>{{Cite journal|last1=Nath|first1=Sunil|last2=Villadsen|first2=John|s2cid=2598635|date=2015-03-01|title=Oxidative phosphorylation revisited|journal=Biotechnology and Bioengineering|language=en|volume=112|issue=3|pages=429–437|doi=10.1002/bit.25492|pmid=25384602|issn=1097-0290}}</ref> and the [[sodium-potassium gradient]] helps [[neural synapse]]s quickly transmit information.{{Citation needed|date=December 2023}}
Electrochemical potential is important in [[electroanalytical chemistry]] and industrial applications such as batteries and fuel cells. It represents one of the many interchangeable forms of [[potential energy]] through which energy may be [[conservation of energy|conserved]].


An electrochemical gradient has two components: a differential concentration of [[electric charge]] across a membrane and a differential concentration of [[chemical species]] across that same membrane. In the former effect, the concentrated charge attracts charges of the opposite sign; in the latter, the concentrated species tends to diffuse across the membrane to an equalize concentrations. The combination of these two phenomena determines the thermodynamically-preferred direction for an [[ion]]'s movement across the membrane.<ref name="LehningerBioChem">{{Cite book|title=Lehninger Principles of Biochemistry|last1=Nelson|first1=David|last2=Cox|first2=Michael|publisher=W.H. Freeman|year=2013|isbn=978-1-4292-3414-6|location=New York}}</ref>{{rp|403}}<ref>{{Cite journal|last1=Yang|first1=Huanghe|last2=Zhang|first2=Guohui|last3=Cui|first3=Jianmin|date=2015-01-01|title=BK channels: multiple sensors, one activation gate|journal= Frontiers in Physiology|volume=6|pages=29|doi=10.3389/fphys.2015.00029|pmc=4319557|pmid=25705194|doi-access=free}}</ref>
In biological processes, the direction an ion moves by [[diffusion]] or [[active transport]] across a membrane is determined by the electrochemical gradient. In [[mitochondria]] and [[chloroplast]]s, proton gradients are used to generate a '''chemiosmotic potential''' that is also known as a '''proton motive force'''. This potential energy is used for the synthesis of ATP by [[oxidative phosphorylation]] or [[photophosphorylation]], respectively.<ref>{{Cite journal|last1=Nath|first1=Sunil|last2=Villadsen|first2=John|s2cid=2598635|date=2015-03-01|title=Oxidative phosphorylation revisited|journal=Biotechnology and Bioengineering|language=en|volume=112|issue=3|pages=429–437|doi=10.1002/bit.25492|pmid=25384602|issn=1097-0290}}</ref>


The combined effect can be quantified as a gradient in the [[thermodynamic]] [[electrochemical potential]]:{{Citation needed|date=December 2023}}<math disply="block">\nabla\overline{\mu}_i = \nabla \mu_i(\vec{r}) + z_i\mathrm{F}\nabla\varphi(\vec{r})\text{,}</math> with <ul>
An electrochemical gradient has two components. First, the electrical component is caused by a charge difference across the lipid membrane. Second, a chemical component is caused by a differential concentration of [[ions]] across the membrane. The combination of these two factors determines the thermodynamically favourable direction for an ion's movement across a membrane.<ref name=":5" /><ref>{{Cite journal|last1=Yang|first1=Huanghe|last2=Zhang|first2=Guohui|last3=Cui|first3=Jianmin|date=2015-01-01|title=BK channels: multiple sensors, one activation gate|journal= Frontiers in Physiology|volume=6|pages=29|doi=10.3389/fphys.2015.00029|pmc=4319557|pmid=25705194}}</ref>
<li>{{math|μ<sub>''i''</sub>}} the chemical potential of the ion species {{mvar|i}}</li>
<li>{{math|''z<sub>i</sub>''}} the charge per ion of the species {{mvar|i}}</li>
<li>{{mvar|F}}, [[Faraday constant]] (the electrochemical potential is implicitly measured on a per-[[mole (chemistry)|mole]] basis)</li>
<li>{{mvar|φ}}, the local [[electric potential]].</li>
</ul> Sometimes, the term "electrochemical potential" is abused to describe the electric potential ''generated'' by an ionic concentration gradient; that is, {{mvar|φ}}.


An electrochemical gradient is analogous to the water [[pressure]] across a [[hydroelectric dam]]. [[Membrane transport protein]]s such as the [[sodium-potassium pump]] within the membrane are equivalent to turbines that convert the water's potential energy to other forms of physical or chemical energy, and the ions that pass through the membrane are equivalent to water that ends up at the bottom of the dam. Also, energy can be used to pump water up into the lake above the dam. In similar manner, chemical energy in cells can be used to create electrochemical gradients.<ref>{{Cite journal|last1=Shattock|first1=Michael J.|last2=Ottolia|first2=Michela|last3=Bers|first3=Donald M.|last4=Blaustein|first4=Mordecai P.|last5=Boguslavskyi|first5=Andrii|last6=Bossuyt|first6=Julie|last7=Bridge|first7=John H. B.|last8=Chen-Izu|first8=Ye|last9=Clancy|first9=Colleen E.|date=2015-03-15|title=Na+/Ca2+ exchange and Na+/K+-ATPase in the heart|journal=The Journal of Physiology|language=en|volume=593|issue=6|pages=1361–1382|doi=10.1113/jphysiol.2014.282319|issn=1469-7793|pmc=4376416|pmid=25772291}}</ref><ref name=":6" />
An electrochemical gradient is analogous to the water [[pressure]] across a [[hydroelectric dam]]. Routes unblocked by the membrane (e.g. [[membrane transport protein]] or [[electrode]]s) correspond to turbines that convert the water's potential energy to other forms of physical or chemical energy, and the ions that pass through the membrane correspond to water traveling into the lower river.{{Tone inline|date=December 2023|reason=Somewhat textbook-like.}} Conversely, energy can be used to [[pumped storage|pump water up into the lake above the dam]], and chemical energy can be used to create electrochemical gradients.<ref>{{Cite journal|last1=Shattock|first1=Michael J.|last2=Ottolia|first2=Michela|last3=Bers|first3=Donald M.|last4=Blaustein|first4=Mordecai P.|last5=Boguslavskyi|first5=Andrii|last6=Bossuyt|first6=Julie|last7=Bridge|first7=John H. B.|last8=Chen-Izu|first8=Ye|last9=Clancy|first9=Colleen E.|date=2015-03-15|title=Na+/Ca2+ exchange and Na+/K+-ATPase in the heart|journal=The Journal of Physiology|language=en|volume=593|issue=6|pages=1361–1382|doi=10.1113/jphysiol.2014.282319|issn=1469-7793|pmc=4376416|pmid=25772291}}</ref><ref name="ReviewWithBadTitle" />


==Chemistry==
==Chemistry==
{{Unsourced|section|date=December 2023}}{{See also|concentration cell|electrode potential|table of standard electrode potentials}}
The term is typically applied in contexts wherein a [[chemical reaction]] is to take place, such as one involving the transfer of an electron at a [[battery (electricity)|battery]] electrode. In a battery, an electrochemical potential arising from the movement of ions balances the reaction energy of the electrodes. The maximum voltage that a battery reaction can produce is sometimes called the '''standard electrochemical potential''' of that reaction (see also [[Electrode potential]] and [[Table of standard electrode potentials]]). In instances pertaining specifically to the movement of electrically charged solutes, the potential is often expressed in units of [[volt]]s. See: [[Concentration cell]].
The term typically applies in [[electrochemistry]], when [[electrical energy]] in the form of an applied voltage is used to modulate the [[Thermodynamic control|thermodynamic favorability]] of a [[chemical reaction]]. In a battery, an electrochemical potential arising from the movement of ions balances the reaction energy of the electrodes. The maximum voltage that a battery reaction can produce is sometimes called the '''standard electrochemical potential''' of that reaction.


==Biological context==
==Biological context==
The generation of a transmembrane electrical potential through ion movement across a cell membrane drives [[biological process]]es like [[nerve]] conduction, [[muscle contraction]], [[hormone]] [[secretion]], and [[Sensory system|sensory]] processes. By convention, a typical animal cell has a transmembrane electrical potential of -50 mV to -70 mV inside the cell relative to the outside.<ref name=":0">{{Cite book|title=Lehninger Principles of Biochemistry|last1=Nelson|first1=David|last2=Cox|first2=Michael|publisher=W.H. Freeman|year=2013|isbn=978-1-4292-3414-6|location=New York|pages=464}}</ref>
The generation of a transmembrane electrical potential through ion movement across a [[cell membrane]] drives [[biological process]]es like [[nerve]] conduction, [[muscle contraction]], [[hormone]] [[secretion]], and [[Sensory system|sensation]]. By convention, physiological voltages are measured [[Gauge transformation|relative]] to the extracellular region; a typical animal cell has an [[transmembrane potential difference|internal electrical potential]] of <!-- {{val|(-70)|-|(-50)|u=mV}} doesn't parenthesize and it looks terrible-->(&minus;70)&ndash;(&minus;50)&nbsp;mV.<ref name="LehningerBioChem" />{{rp|464}}


Electrochemical gradients also play a role in establishing proton gradients in oxidative phosphorylation in mitochondria. The final step of [[cellular respiration]] is the [[electron transport chain]]. Four complexes embedded in the inner membrane of the mitochondrion make up the electron transport chain. However, only complexes I, III, and IV pump protons from the [[Mitochondrial matrix|matrix]] to the [[Mitochondrial intermembrane space|intermembrane space]] (IMS). In total, there are ten protons translocated from the matrix to the IMS which generates an electrochemical potential of more than 200mV. This drives the flux of protons back into the matrix through [[ATP synthase]] which produces [[Adenosine triphosphate|ATP]] by adding an inorganic [[phosphate]] to [[Adenosine diphosphate|ADP]].<ref>{{Cite journal|last1=Poburko|first1=Damon|last2=Demaurex|first2=Nicolas|date=2012-04-24|title=Regulation of the mitochondrial proton gradient by cytosolic Ca2+ signals|journal=Pflügers Archiv: European Journal of Physiology|language=en|volume=464|issue=1|pages=19–26|doi=10.1007/s00424-012-1106-y|pmid=22526460|s2cid=18133149|issn=0031-6768|url=http://doc.rero.ch/record/321929/files/424_2012_Article_1106.pdf}}</ref> Thus, generation of a proton electrochemical gradient is crucial for energy production in mitochondria.<ref>{{Cite book|title=Lehninger Principles of Biochemistry|last1=Nelson|first1=David|last2=Cox|first2=Michael|publisher=W.H. Freeman|year=2013|isbn=978-1-4292-3414-6|location=New York|pages=743–745}}</ref> The total equation for the electron transport chain is:
An electrochemical gradient is essential to [[mitochondrial]] [[oxidative phosphorylation]]. The final step of [[cellular respiration]] is the [[electron transport chain]], composed of four complexes embedded in the inner mitochondrial membrane. Complexes I, III, and IV pump protons from the [[Mitochondrial matrix|matrix]] to the [[Mitochondrial intermembrane space|intermembrane space]] (IMS); for every [[electron pair]] entering the chain, ten protons translocate into the IMS. The result is an electric potential of more than {{val|200|u=mV}}. The resulting flux of protons back into the matrix powers the efforts of [[ATP synthase]] to combine inorganic [[phosphate]] and [[Adenosine diphosphate|ADP]].<ref>{{Cite journal|last1=Poburko|first1=Damon|last2=Demaurex|first2=Nicolas|date=2012-04-24|title=Regulation of the mitochondrial proton gradient by cytosolic Ca2+ signals|journal=Pflügers Archiv: European Journal of Physiology|language=en|volume=464|issue=1|pages=19–26|doi=10.1007/s00424-012-1106-y|pmid=22526460|s2cid=18133149|issn=0031-6768|url=http://doc.rero.ch/record/321929/files/424_2012_Article_1106.pdf}}</ref><ref name=LehningerBioChem />{{rp|743–745}}


Similar to the electron transport chain, the [[light-dependent reactions]] of photosynthesis pump protons into the [[thylakoid]] [[Lumen (anatomy)|lumen]] of chloroplasts to drive the synthesis of ATP. The proton gradient can be generated through either noncyclic or cyclic photophosphorylation. Of the proteins that participate in noncyclic photophosphorylation, [[photosystem II]] (PSII), [[Plastoquinone|plastiquinone]], and [[Cytochrome b6f complex|cytochrome b<sub>6</sub>f complex]] directly contribute to generating the proton gradient. For each four photons absorbed by PSII, eight protons are pumped into the lumen.<ref name=LehningerBioChem />{{rp|769–770}}
NADH + 11 H<sup>+</sup><sub>(matrix)</sub> + 1/2 O<sub>2</sub> → NAD<sup>+</sup> + 10 H<sup>+</sup><sub>(IMS)</sub> + H<sub>2</sub>O.<ref>{{Cite book|title=Lehninger Principles of Biochemistry|last1=Nelson|first1=David|last2=Cox|first2=Michael|publisher=W.H. Freeman|year=2013|isbn=978-1-4292-3414-6|location=New York|pages=744}}</ref>


Several other transporters and ion channels play a role in generating a proton electrochemical gradient. One is TPK<sub>3</sub>, a [[potassium channel]] that is activated by Ca<sup>2+</sup> and conducts K<sup>+</sup> from the thylakoid lumen to the [[Stroma (fluid)|stroma]], which helps establish the [[electric field]]. On the other hand, the electro-neutral K<sup>+</sup> efflux [[antiporter]] (KEA<sub>3</sub>) transports K<sup>+</sup> into the thylakoid lumen and H<sup>+</sup> into the stroma, which helps establish the [[pH]] gradient.<ref>{{Cite journal|last1=Höhner|first1=Ricarda|last2=Aboukila|first2=Ali|last3=Kunz|first3=Hans-Henning|last4=Venema|first4=Kees|date=2016-01-01|title= Proton Gradients and Proton-Dependent Transport Processes in the Chloroplast|journal= Frontiers in Plant Science|pages=218|doi=10.3389/fpls.2016.00218|pmc=4770017|pmid=26973667|volume=7|doi-access=free}}</ref>
Similar to the electron transport chain, the [[light-dependent reactions]] of photosynthesis pump protons into the [[thylakoid]] [[Lumen (anatomy)|lumen]] of chloroplasts to drive the synthesis of ATP by ATP synthase. The proton gradient can be generated through either noncyclic or cyclic photophosphorylation. Of the proteins that participate in noncyclic photophosphorylation, [[photosystem II]] (PSII), [[Plastoquinone|plastiquinone]], and [[Cytochrome b6f complex|cytochrome b<sub>6</sub>f complex]] directly contribute to generating the proton gradient. For each four photons absorbed by PSII, eight protons are pumped into the lumen.<ref>{{Cite book|title=Lehninger Principles of Biochemistry|last1=Nelson|first1=David|last2=Cox|first2=Michael|publisher=W.H. Freeman|year=2013|isbn=978-1-4292-3414-6|location=New York|pages=769–770}}</ref> The total equation for photophosphorylation is:

2 NADP<sup>+</sup> + 6 H<sup>+</sup><sub>(stroma)</sub> + 2 H<sub>2</sub>O → 2 NADPH + 8 H<sup>+</sup><sub>(lumen)</sub> + O<sub>2</sub>.<ref>{{Cite book|title=Lehninger Principles of Biochemistry|last1=Nelson|first1=David|last2=Cox|first2=Michael|publisher=W.H. Freeman|year=2013|isbn=978-1-4292-3414-6|location=New York|pages=770}}</ref>

Several other transporters and ion channels play a role in generating a proton electrochemical gradient. One is TPK<sub>3</sub>, a [[potassium channel]] that is activated by Ca<sup>2+</sup> and conducts K<sup>+</sup> from the thylakoid lumen to the [[Stroma (fluid)|stroma]] which helps establish the [[pH]] gradient. On the other hand, the electro-neutral K<sup>+</sup> efflux [[antiporter]] (KEA<sub>3</sub>) transports K<sup>+</sup> into the thylakoid lumen and H<sup>+</sup> into the stroma which helps establish the [[electric field]].<ref>{{Cite journal|last1=Höhner|first1=Ricarda|last2=Aboukila|first2=Ali|last3=Kunz|first3=Hans-Henning|last4=Venema|first4=Kees|date=2016-01-01|title= Proton Gradients and Proton-Dependent Transport Processes in the Chloroplast|journal= Frontiers in Plant Science|pages=218|doi=10.3389/fpls.2016.00218|pmc=4770017|pmid=26973667|volume=7}}</ref>


==Ion gradients==
==Ion gradients==
[[File:Scheme sodium-potassium pump-en.svg|thumb|Diagram of the Na<sup>+</sup>-K<sup>+</sup>-ATPase.]]
[[File:Scheme sodium-potassium pump-en.svg|thumb|Diagram of the Na<sup>+</sup>-K<sup>+</sup>-ATPase.]]
Since the ions are charged, they cannot pass through cellular membranes via simple diffusion. Two different mechanisms can transport the ions across the membrane: [[Active transport|active]] or [[Passive transport|passive]] transport.{{Citation needed|date=December 2023}}
Since the ions are charged, they cannot pass through the membrane via simple diffusion. Two different mechanisms can transport the ions across the membrane: [[Active transport|active]] or [[Passive transport|passive]] transport. An example of active transport of ions is the [[Na+/K+-ATPase|Na<sup>+</sup>-K<sup>+</sup>-ATPase]] (NKA). NKA catalyzes the [[hydrolysis]] of ATP into ADP and an inorganic phosphate and for every molecule of ATP hydrolized, three Na<sup>+</sup> are transported outside and two K<sup>+</sup> are transported inside the cell. This makes the inside of the cell more negative than the outside and more specifically generates a membrane potential ''V''<sub>membrane</sub> of about -60mV.<ref name=":6">{{Cite journal|last1=Aperia|first1=Anita|last2=Akkuratov|first2=Evgeny E.|last3=Fontana|first3=Jacopo Maria|last4=Brismar|first4=Hjalmar|date=2016-04-01|title=Na+-K+-ATPase, a new class of plasma membrane receptors|journal=American Journal of Physiology. Cell Physiology|language=en|volume=310|issue=7|pages=C491–C495|doi=10.1152/ajpcell.00359.2015|issn=0363-6143|pmid=26791490|url=https://zenodo.org/record/1065636|doi-access=free}}</ref> An example of passive transport is ion fluxes through Na<sup>+</sup>, K<sup>+</sup>, Ca<sup>2+</sup>, and Cl<sup>−</sup> channels. These ions tend to move down their concentration gradient. For example, since there is a high concentration of Na<sup>+</sup> outside the cell, Na<sup>+</sup> will flow through the Na<sup>+</sup> channel into the cell. Since the electric potential inside the cell is negative, the influx of a positive ion depolarizes the membrane which brings the transmembrane electric potential closer to zero. However, Na<sup>+</sup> will continue moving down its concentration gradient as long as the effect of the chemical gradient is greater than the effect of the electrical gradient. Once the effect of both gradients are equal (for Na<sup>+</sup> this at a membrane potential of about +70mV), the influx of Na<sup>+</sup> stops because the [[Reversal potential|driving force]] (ΔG) is zero. The equation for the driving force is:<ref>{{Cite book|title=Lehninger Principles of Biochemistry|last1=Nelson|first1=David|last2=Cox|first2=Michael|publisher=W.H. Freeman|year=2013|isbn=978-1-4292-3414-6|location=New York|pages=464–465}}</ref><ref name=":2">{{Cite journal|last=Eisenberg|first=Bob|date=2013-05-07|title=Interacting Ions in Biophysics: Real is not Ideal|url=http://www.cell.com/biophysj/abstract/S0006-3495%2813%2900388-3|journal=Biophysical Journal|volume=104|issue=9|pages=1849–1866|doi=10.1016/j.bpj.2013.03.049|pmid=23663828|pmc=3647150|arxiv=1305.2086|bibcode=2013BpJ...104.1849E}}</ref>


An example of active transport of ions is the [[Na+/K+-ATPase|Na<sup>+</sup>-K<sup>+</sup>-ATPase]] (NKA). NKA is powered by the [[hydrolysis]] of ATP into ADP and an inorganic phosphate; for every molecule of ATP hydrolized, three Na<sup>+</sup> are transported outside and two K<sup>+</sup> are transported inside the cell. This makes the inside of the cell more negative than the outside and more specifically generates a membrane potential ''V''<sub>membrane</sub> of about {{val|-60|u=mV}}.<ref name="ReviewWithBadTitle">{{Cite journal|last1=Aperia|first1=Anita|last2=Akkuratov|first2=Evgeny E.|last3=Fontana|first3=Jacopo Maria|last4=Brismar|first4=Hjalmar|date=2016-04-01|title=Na+-K+-ATPase, a new class of plasma membrane receptors|journal=American Journal of Physiology. Cell Physiology|language=en|volume=310|issue=7|pages=C491–C495|doi=10.1152/ajpcell.00359.2015|issn=0363-6143|pmid=26791490|url=https://zenodo.org/record/1065636|doi-access=free}}</ref>
<math> \Delta G = RT \ln \frac{c_{\rm in}}{c_{\rm out}} + zFV_{\rm membrane}</math><ref name=":0" />


An example of passive transport is ion fluxes through Na<sup>+</sup>, K<sup>+</sup>, Ca<sup>2+</sup>, and Cl<sup>−</sup> channels. Unlike active transport, passive transport is powered by the [[arithmetic sum]] of [[osmosis]] (a concentration gradient) and an [[electric field]] (the transmembrane potential). Formally, the [[Mole (chemistry)|molar]] [[Gibbs free energy]] change associated with successful transport is{{Citation needed|date=December 2023}} <math display=block> \Delta G = RT\ln{\!\left(\frac{c_{\rm in}}{c_{\rm out}}\right)} + (Fz)V_{\rm membrane}</math> where {{mvar|R}} represents the [[gas constant]], {{mvar|T}} represents [[Thermodynamic temperature|absolute temperature]], {{mvar|z}} is the charge per ion, and {{mvar|F}} represents the [[Faraday constant]].<ref name="LehningerBioChem" />{{rp|464–465}}
In this equation, ''R'' represents the [[gas constant]], ''T'' represents [[Thermodynamic temperature|absolute temperature]], ''z'' is the ionic charge, and ''F'' represents the [[Faraday constant]].<ref name=":3">{{Cite book|title=Lehninger Principles of Biochemistry|last1=Nelson|first1=David|last2=Cox|first2=Michael|publisher=W.H. Freeman|year=2013|isbn=978-1-4292-3414-6|location=New York|pages=465}}</ref>

In the example of Na<sup>+</sup>, both terms tend to support transport: the negative electric potential inside the cell attracts the positive ion and since Na<sup>+</sup> is concentrated outside the cell, osmosis supports diffusion through the Na<sup>+</sup> channel into the cell. In the case of K<sup>+</sup>, the effect of osmosis is reversed: although external ions are attracted by the negative intracellular potential, entropy seeks to diffuse the ions already concentrated inside the cell. The converse phenomenon (osmosis supports transport, electric potential opposes it) can be achieved for Na<sup>+</sup> in cells with abnormal transmembrane potentials: at {{val|+70|u=mV}}, the Na<sup>+</sup> influx halts; at higher potentials, it becomes an efflux.{{Citation needed|date=December 2023}}


Cellular ion concentrations are given in the table below. X- represents proteins with a net negative charge.
{| class="wikitable"
{| class="wikitable"
|+
|+
Cellular ion concentrations ([[Molar concentration|millimolar]])<ref>{{Cite web|url=http://book.bionumbers.org/what-are-the-concentrations-of-different-ions-in-cells/|title=» What are the concentrations of different ions in cells?|last=Philips|first=Ron Milo & Ron|language=en|access-date=2019-06-07}}</ref><ref>{{Cite web|url=https://www.ncbi.nlm.nih.gov/books/NBK21627/table/A4057/|title=Table 15-1, Typical Ion Concentrations in Invertebrates and Vertebrates|last1=Lodish|first1=Harvey|last2=Berk|first2=Arnold|date=2000|website=www.ncbi.nlm.nih.gov|language=en|access-date=2019-06-07|last3=Zipursky|first3=S. Lawrence|last4=Matsudaira|first4=Paul|last5=Baltimore|first5=David|last6=Darnell|first6=James}}</ref><ref>{{Cite web|url=http://www.chm.bris.ac.uk/webprojects2001/riis/ionconc.htm|title=The following table gives an idea of the intra and extra cellular ion concentrations in a squid axon and a mammalian cell|website=www.chm.bris.ac.uk|access-date=2019-06-07}}</ref><ref name="diem">{{cite book|title=Scientific Tables|vauthors=Diem K, Lenter C|publisher=Ciba-Geigy Limited|isbn=978-3-9801244-0-9|edition=Seventh|volume=565|location=Basel|pages=653–654}}</ref>
Common cellular ion concentrations ([[Molar concentration|millimolar]])<ref>{{Cite web|url=http://book.bionumbers.org/what-are-the-concentrations-of-different-ions-in-cells/|title=» What are the concentrations of different ions in cells?|last=Philips|first=Ron Milo & Ron|language=en|access-date=2019-06-07}}</ref><ref>{{Cite web|url=https://www.ncbi.nlm.nih.gov/books/NBK21627/table/A4057/|title=Table 15-1, Typical Ion Concentrations in Invertebrates and Vertebrates|last1=Lodish|first1=Harvey|last2=Berk|first2=Arnold|date=2000|website=www.ncbi.nlm.nih.gov|language=en|access-date=2019-06-07|last3=Zipursky|first3=S. Lawrence|last4=Matsudaira|first4=Paul|last5=Baltimore|first5=David|last6=Darnell|first6=James}}</ref><ref>{{Cite web|url=http://www.chm.bris.ac.uk/webprojects2001/riis/ionconc.htm|title=The following table gives an idea of the intra and extra cellular ion concentrations in a squid axon and a mammalian cell|website=www.chm.bris.ac.uk|access-date=2019-06-07}}</ref><ref name="diem">{{cite book|title=Scientific Tables|vauthors=Diem K, Lenter C|publisher=Ciba-Geigy Limited|isbn=978-3-9801244-0-9|edition=Seventh|volume=565|location=Basel|pages=653–654}}</ref>
! rowspan="2" |Ion
! rowspan="2" |Ion
! colspan="2" |[[Mammal]]
! colspan="2" |[[Mammal]]
Line 61: Line 63:
!Blood
!Blood
|-
|-
|[[Potassium in biology|K<sup>+</sup>]]
|[[Potassium in biology|K<sup>+</sup>]] || 100 - 140 || 4-5 || 400 || 10 - 20 || 300 || 30 - 300 || 10
|100 - 140
|4-5
|400
|10 - 20
|300
|30 - 300
|10
|-
|-
|[[Sodium in biology|Na<sup>+</sup>]]
|[[Sodium in biology|Na<sup>+</sup>]] || 5-15 || 145 || 50 || 440 || 30 || 10 || 500
|5-15
|145
|50
|440
|30
|10
|500
|-
|-
|[[Magnesium in biology|Mg<sup>2+</sup>]]
|[[Magnesium in biology|Mg<sup>2+</sup>]] || 10 {{efn|name=bound|Bound}}<br>0.5 - 0.8 {{efn|name=free|Free}} || 1 - 1.5 || || || 50 || 30 - 100 {{efn|name=bound}}<br>0.01 - 1 {{efn|name=free}} || 50
|10 {{efn|name=bound|Bound}}<br>0.5 - 0.8 {{efn|name=free|Free}}
|1 - 1.5
|
|
|50
|30 - 100 {{efn|name=bound}}<br>0.01 - 1 {{efn|name=free}}
|50
|-
|-
|[[Calcium in biology|Ca<sup>2+</sup>]] || 10<sup>−4</sup> || 2.2 - 2.6 {{efn|name=total|Total}}<br>1.3 - 1.5 {{efn|name=ionised|Ionised}} || 10<sup>−4</sup> - 3×10<sup>−4</sup> || 10 || 2 || 3 {{efn|name=bound}}<br>10<sup>−4</sup> {{efn|name=free}} || 10
|[[Calcium in biology|Ca<sup>2+</sup>]]
|10<sup>−4</sup>
|2.2 - 2.6 {{efn|name=total|Total}}<br>1.3 - 1.5 {{efn|name=ionised|Ionised}}
|10<sup>−4</sup> - 3×10<sup>−4</sup>
|10
|2
|3 {{efn|name=bound}}<br>10<sup>−4</sup> {{efn|name=free}}
|10
|-
|-
|[[Chloride|Cl<sup>−</sup>]]
|[[Chloride|Cl<sup>−</sup>]] || 4 || 110 || 40 - 150 || 560
| colspan="2" |10 - 200 {{efn|Medium dependent}} || 500
|4
|110
|40 - 150
|560
| colspan="2" |10 - 200 {{efn|Medium dependent}}
|500
|-
|-
|[[Protein|X<sup>−</sup>]]
|[[Protein|X<sup>−</sup> (negatively charged proteins)]] || 138 || 9 || 300 - 400 || 5-10 || || ||
|138
|9
|300 - 400
|5-10
|
|
|
|-
|-
|[[Bicarbonate|HCO<sub>3</sub><sup>−</sup>]]
|[[Bicarbonate|HCO<sub>3</sub><sup>−</sup>]] || 12 || 29 || || || || ||
|12
|29
|
|
|
|
|
|-
|-
|pH || 7.1 - 7.3<ref name=":7">{{Citation|last1=Spitzer|first1=Kenneth W.|title=Regulation of Intracellular pH in Mammalian Cells|date=2003|work=The Sodium-Hydrogen Exchanger: From Molecule to its Role in Disease|pages=1–15|editor-last=Karmazyn|editor-first=Morris|publisher=Springer US|language=en|doi=10.1007/978-1-4615-0427-6_1|isbn=9781461504276|last2=Vaughan-Jones|first2=Richard D.|editor2-last=Avkiran|editor2-first=Metin|editor3-last=Fliegel|editor3-first=Larry}}</ref> || 7.35 to 7.45 <ref name=":7" /> (normal arterial blood pH)<br />6.9 - 7.8 <ref name=":7" /> (overall range) || || || ||7.2 - 7.8<ref>{{Cite journal|last1=Slonczewski|first1=Joan L.|last2=Wilks|first2=Jessica C.|date=2007-08-01|title=pH of the Cytoplasm and Periplasm of Escherichia coli: Rapid Measurement by Green Fluorescent Protein Fluorimetry|journal=Journal of Bacteriology|language=en|volume=189|issue=15|pages=5601–5607|doi=10.1128/JB.00615-07|issn=0021-9193|pmid=17545292|pmc=1951819}}</ref>
|pH
|7.1 - 7.3<ref name=":7">{{Citation|last1=Spitzer|first1=Kenneth W.|title=Regulation of Intracellular pH in Mammalian Cells|date=2003|work=The Sodium-Hydrogen Exchanger: From Molecule to its Role in Disease|pages=1–15|editor-last=Karmazyn|editor-first=Morris|publisher=Springer US|language=en|doi=10.1007/978-1-4615-0427-6_1|isbn=9781461504276|last2=Vaughan-Jones|first2=Richard D.|editor2-last=Avkiran|editor2-first=Metin|editor3-last=Fliegel|editor3-first=Larry}}</ref>
|7.35 to 7.45 <ref name=":7" /> (normal arterial blood pH)
6.9 - 7.8 <ref name=":7" /> (overall range)
|
|
|
|7.2 - 7.8<ref>{{Cite journal|last1=Slonczewski|first1=Joan L.|last2=Wilks|first2=Jessica C.|date=2007-08-01|title=pH of the Cytoplasm and Periplasm of Escherichia coli: Rapid Measurement by Green Fluorescent Protein Fluorimetry|journal=Journal of Bacteriology|language=en|volume=189|issue=15|pages=5601–5607|doi=10.1128/JB.00615-07|issn=0021-9193|pmid=17545292|pmc=1951819}}</ref>
|8.1 - 8.2<ref>{{Cite journal|url=https://www.scientificamerican.com/article/rising-acidity-in-the-ocean/|title=Rising Acidity in the Ocean: The Other CO2 Problem|last=Brewer|first=Peter G.|date=September 1, 2008|doi=10.1038/scientificamericanearth0908-22}}</ref>
|8.1 - 8.2<ref>{{Cite journal|url=https://www.scientificamerican.com/article/rising-acidity-in-the-ocean/|title=Rising Acidity in the Ocean: The Other CO2 Problem|last=Brewer|first=Peter G.|date=September 1, 2008|doi=10.1038/scientificamericanearth0908-22}}</ref>
|}
|}
Line 137: Line 85:


==Proton gradients==
==Proton gradients==
Proton gradients in particular are important in many types of cells as a form of energy storage. The gradient is usually used to drive ATP synthase, [[Flagellum|flagellar]] rotation, or transport of [[metabolite]]s.<ref name=":4">{{Cite journal|last1=Gunner|first1=M. R.|last2=Amin|first2=Muhamed|last3=Zhu|first3=Xuyu|last4=Lu|first4=Jianxun|date=2013-08-01|title=Molecular mechanisms for generating transmembrane proton gradients|journal=Biochimica et Biophysica Acta (BBA) - Bioenergetics|series=Metals in Bioenergetics and Biomimetics Systems|volume=1827|issue=8–9|pages=892–913|doi=10.1016/j.bbabio.2013.03.001|pmc=3714358|pmid=23507617}}</ref> This section will focus on three processes that help establish proton gradients in their respective cells: [[bacteriorhodopsin]] and noncyclic photophosphorylation and oxidative phosphorylation.
Proton gradients in particular are important in many types of cells as a form of energy storage. The gradient is usually used to drive ATP synthase, [[Flagellum|flagellar]] rotation, or [[metabolite]] transport.<ref name="H+PumpReview">{{Cite journal|last1=Gunner|first1=M. R.|last2=Amin|first2=Muhamed|last3=Zhu|first3=Xuyu|last4=Lu|first4=Jianxun|date=2013-08-01|title=Molecular mechanisms for generating transmembrane proton gradients|journal=Biochimica et Biophysica Acta (BBA) - Bioenergetics|series=Metals in Bioenergetics and Biomimetics Systems|volume=1827|issue=8–9|pages=892–913|doi=10.1016/j.bbabio.2013.03.001|pmc=3714358|pmid=23507617}}</ref> This section will focus on three processes that help establish proton gradients in their respective cells: [[bacteriorhodopsin]] and noncyclic photophosphorylation and oxidative phosphorylation.{{Citation needed|date=December 2023}}


=== Bacteriorhodopsin ===
=== Bacteriorhodopsin ===
[[File:Bacteriorhodopsin retinal.png|thumb|Diagram of the conformational shift in retinal that initiates proton pumping in bacteriorhodopsin.]]
[[File:Bacteriorhodopsin retinal.png|thumb|Diagram of the conformational shift in retinal that initiates proton pumping in bacteriorhodopsin.]]
The way bacteriorhodopsin generates a proton gradient in [[Archaea]] is through a [[proton pump]]. The proton pump relies on proton carriers to drive protons from the side of the membrane with a low H<sup>+</sup> concentration to the side of the membrane with a high H<sup>+</sup> concentration. In bacteriorhodopsin, the proton pump is activated by absorption of [[photon]]s of 568&nbsp;nm [[wavelength]] which leads to [[isomerization]] of the [[Schiff base]] (SB) in [[retinal]] forming the K state. This moves SB away from Asp85 and Asp212, causing H<sup>+</sup> transfer from the SB to Asp85 forming the M1 state. The protein then shifts to the M2 state by separating Glu204 from Glu194 which releases a proton from Glu204 into the external medium. The SB is [[Protonation|reprotonated]] by Asp96 which forms the N state. It is important that the second proton comes from Asp96 since its [[Deprotonation|deprotonated]] state is unstable and rapidly reprotonated with a proton from the [[cytosol]]. The protonation of Asp85 and Asp96 causing re-isomerization of the SB forming the O state. Finally, bacteriorhodopsin returns to its resting state when Asp85 releases its proton to Glu204.<ref name=":4" /><ref>{{Cite journal|last1=Wickstrand|first1=Cecilia|last2=Dods|first2=Robert|last3=Royant|first3=Antoine|last4=Neutze|first4=Richard|date=2015-03-01|title=Bacteriorhodopsin: Would the real structural intermediates please stand up?|journal=Biochimica et Biophysica Acta (BBA) - General Subjects|series=Structural biochemistry and biophysics of membrane proteins|volume=1850|issue=3|pages=536–553|doi=10.1016/j.bbagen.2014.05.021|pmid=24918316|doi-access=free}}</ref>
The way [[bacteriorhodopsin]] generates a proton gradient in [[Archaea]] is through a [[proton pump]]. The proton pump relies on proton carriers to drive protons from the side of the membrane with a low H<sup>+</sup> concentration to the side of the membrane with a high H<sup>+</sup> concentration. In bacteriorhodopsin, the proton pump is activated by absorption of [[photon]]s of 568nm [[wavelength]], which leads to [[isomerization]] of the [[Schiff base]] (SB) in [[retinal]] forming the K state. This moves SB away from Asp85 and Asp212, causing H<sup>+</sup> transfer from the SB to Asp85 forming the M1 state. The protein then shifts to the M2 state by separating Glu204 from Glu194 which releases a proton from Glu204 into the external medium. The SB is [[Protonation|reprotonated]] by Asp96 which forms the N state. It is important that the second proton comes from Asp96 since its [[Deprotonation|deprotonated]] state is unstable and rapidly reprotonated with a proton from the [[cytosol]]. The protonation of Asp85 and Asp96 causes re-isomerization of the SB, forming the O state. Finally, bacteriorhodopsin returns to its resting state when Asp85 releases its proton to Glu204.<ref name="H+PumpReview" /><ref>{{Cite journal|last1=Wickstrand|first1=Cecilia|last2=Dods|first2=Robert|last3=Royant|first3=Antoine|last4=Neutze|first4=Richard|date=2015-03-01|title=Bacteriorhodopsin: Would the real structural intermediates please stand up?|journal=Biochimica et Biophysica Acta (BBA) - General Subjects|series=Structural biochemistry and biophysics of membrane proteins|volume=1850|issue=3|pages=536–553|doi=10.1016/j.bbagen.2014.05.021|pmid=24918316|doi-access=free}}</ref>


=== Photophosphorylation ===
=== Photophosphorylation ===
[[File:Cyclic Photophosphorylation.svg|thumb|Simplified diagram of photophosphorylation.]]
[[File:Cyclic Photophosphorylation.svg|thumb|Simplified diagram of photophosphorylation.]]
PSII also relies on [[light]] to drive the formation of proton gradients in chloroplasts, however PSII utilizes vectorial redox chemistry to achieve this goal. Rather than physically transporting protons through the protein, reactions requiring the binding of protons will occur on the extracellular side while reactions requiring the release of protons will occur on the intracellular side. Absorption of photons of 680&nbsp;nm wavelength is used to excite two electrons in [[P680|P<sub>680</sub>]] to a higher [[energy level]]. These higher energy electrons are transferred to protein-bound plastoquinone (PQ<sub>A</sub>) and then to unbound plastoquinone (PQ<sub>B</sub>). This reduces plastoquinone (PQ) to plastoquinol (PQH<sub>2</sub>) which is released from PSII after gaining two protons from the stroma. The electrons in P<sub>680</sub> are replenished by oxidizing [[water]] through the [[oxygen-evolving complex]] (OEC). This results in release of O<sub>2</sub> and H<sup>+</sup> into the lumen.<ref name=":4" /> The total reaction is shown:
PSII also relies on [[light]] to drive the formation of proton gradients in chloroplasts, however, PSII utilizes vectorial redox chemistry to achieve this goal. Rather than physically transporting protons through the protein, reactions requiring the binding of protons will occur on the extracellular side while reactions requiring the release of protons will occur on the intracellular side. Absorption of photons of 680nm wavelength is used to excite two electrons in [[P680|P<sub>680</sub>]] to a higher [[energy level]]. These higher energy electrons are transferred to protein-bound [[plastoquinone]] (PQ<sub>A</sub>) and then to unbound plastoquinone (PQ<sub>B</sub>). This reduces plastoquinone (PQ) to plastoquinol (PQH<sub>2</sub>) which is released from PSII after gaining two protons from the stroma. The electrons in P<sub>680</sub> are replenished by oxidizing [[water]] through the [[oxygen-evolving complex]] (OEC). This results in release of O<sub>2</sub> and H<sup>+</sup> into the lumen, for a total reaction of<ref name="H+PumpReview" />
<math chem display=block>4h\nu+2\ce{H2O}+2\ce{PQ}+4\ce{H+}(\text{stroma})\longrightarrow\ce{O2}+2\ce{PQH2}+4\ce{H+}(\text{lumen})</math>


After being released from PSII, PQH<sub>2</sub> travels to the [[Cytochrome b6f complex|cytochrome b<sub>6</sub>f complex]], which then transfers two electrons from PQH<sub>2</sub> to [[plastocyanin]] in two separate reactions. The process that occurs is similar to the Q-cycle in Complex III of the electron transport chain. In the first reaction, PQH<sub>2</sub> binds to the complex on the lumen side and one electron is transferred to the [[Iron–sulfur protein|iron-sulfur center]] which then transfers it to [[cytochrome f]] which then transfers it to plastocyanin. The second electron is transferred to [[Heme B|heme b<sub>L</sub>]] which then transfers it to heme b<sub>H</sub> which then transfers it to PQ. In the second reaction, a second PQH<sub>2</sub> gets oxidized, adding an electron to another plastocyanin and PQ. Both reactions together transfer four protons into the lumen.<ref name=LehningerBioChem />{{rp|782–783}}<ref>{{Cite journal|last1=Schöttler|first1=Mark Aurel|last2=Tóth|first2=Szilvia Z.|last3=Boulouis|first3=Alix|last4=Kahlau|first4=Sabine|date=2015-05-01|title=Photosynthetic complex stoichiometry dynamics in higher plants: biogenesis, function, and turnover of ATP synthase and the cytochrome b 6 f complex|journal=Journal of Experimental Botany|language=en|volume=66|issue=9|pages=2373–2400|doi=10.1093/jxb/eru495|issn=0022-0957|pmid=25540437|doi-access=free}}</ref>
<math>4\ photons(680nm)\ +\ 2H_2O\ +\ 2PQ\ +\ 4H^+(stroma)\longrightarrow O_2\ +\ 2PQH_2\ +\ 4H^+(lumen)</math><ref name=":4" />

After being released from PSII, PQH<sub>2</sub> travels to the cytochrome b<sub>6</sub>f complex which then transfers two electrons from PQH<sub>2</sub> to [[plastocyanin]] in two separate reactions. The process that occurs is similar to the Q-cycle in Complex III of the electron transport chain. In the first reaction, PQH<sub>2</sub> binds to the complex on the lumen side and one electron is transferred to the [[Iron–sulfur protein|iron-sulfur center]] which then transfers it to [[cytochrome f]] which then transfers it to plastocyanin. The second electron is transferred to [[Heme B|heme b<sub>L</sub>]] which then transfers it to heme b<sub>H</sub> which then transfers it to PQ. In the second reaction, a second PQH<sub>2</sub> gets oxidized, adding an electron to another plastocyanin and PQ. Both reactions together transfer four protons into the lumen.<ref>{{Cite book|title=Lehninger Principles of Biochemistry|last1=Nelson|first1=David|last2=Cox|first2=Michael|publisher=W.H. Freeman|year=2013|isbn=978-1-4292-3414-6|location=New York|pages=782–783}}</ref><ref>{{Cite journal|last1=Schöttler|first1=Mark Aurel|last2=Tóth|first2=Szilvia Z.|last3=Boulouis|first3=Alix|last4=Kahlau|first4=Sabine|date=2015-05-01|title=Photosynthetic complex stoichiometry dynamics in higher plants: biogenesis, function, and turnover of ATP synthase and the cytochrome b 6 f complex|journal=Journal of Experimental Botany|language=en|volume=66|issue=9|pages=2373–2400|doi=10.1093/jxb/eru495|issn=0022-0957|pmid=25540437|doi-access=free}}</ref>


=== Oxidative phosphorylation ===
=== Oxidative phosphorylation ===
[[File:ETC electron transport chain.svg|thumb|Detailed diagram of the electron transport chain in mitochondria.]]
[[File:ETC electron transport chain.svg|thumb|Detailed diagram of the electron transport chain in mitochondria.]]
In the electron transport chain, [[complex I]] (CI) [[Catalysis|catalyzes]] the [[Redox|reduction]] of [[Coenzyme Q10|ubiquinone]] (UQ) to [[ubiquinol]] (UQH<sub>2</sub>) by the transfer of two [[electron]]s from reduced [[nicotinamide adenine dinucleotide]] (NADH) which translocates four protons from the mitochondrial matrix to the IMS:<ref name=":1">{{Cite journal|last1=Sun|first1=Fei|last2=Zhou|first2=Qiangjun|last3=Pang|first3=Xiaoyun|last4=Xu|first4=Yingzhi|last5=Rao|first5=Zihe|date=2013-08-01|title=Revealing various coupling of electron transfer and proton pumping in mitochondrial respiratory chain|journal=Current Opinion in Structural Biology|volume=23|issue=4|pages=526–538|doi=10.1016/j.sbi.2013.06.013|pmid=23867107}}</ref>
In the electron transport chain, [[complex I]] (CI) [[Catalysis|catalyzes]] the [[Redox|reduction]] of [[Coenzyme Q10|ubiquinone]] (UQ) to [[ubiquinol]] (UQH<sub>2</sub>) by the transfer of two [[electron]]s from reduced [[nicotinamide adenine dinucleotide]] (NADH) which translocates four protons from the mitochondrial matrix to the IMS:<ref name="PumpStructures">{{Cite journal|last1=Sun|first1=Fei|last2=Zhou|first2=Qiangjun|last3=Pang|first3=Xiaoyun|last4=Xu|first4=Yingzhi|last5=Rao|first5=Zihe|date=2013-08-01|title=Revealing various coupling of electron transfer and proton pumping in mitochondrial respiratory chain|journal=Current Opinion in Structural Biology|volume=23|issue=4|pages=526–538|doi=10.1016/j.sbi.2013.06.013|pmid=23867107}}</ref>
<math chem display=block>\ce{NADH} + \ce{H^+} + \ce{UQ} + 4\underbrace{\ce{H^+}}_{\mathrm{matrix}} \longrightarrow \ce{NAD^+} + \ce{UQH_2} + 4\underbrace{\ce{H^+}}_{\mathrm{IMS}}</math>

<math chem>\ce{NADH} + \ce{H^+} + \ce{UQ} + 4\underbrace{\ce{H^+}}_{\mathrm{matrix}} \longrightarrow \ce{NAD^+} + \ce{UQH_2} + 4\underbrace{\ce{H^+}}_{\mathrm{IMS}}</math><ref name=":1" />

[[Complex III]] (CIII) catalyzes the [[Q cycle|Q-cycle]]. The first step involving the transfer of two electrons from the UQH<sub>2</sub> reduced by CI to two molecules of oxidized [[cytochrome c]] at the Q<sub>o</sub> site. In the second step, two more electrons reduce UQ to UQH<sub>2</sub> at the Q<sub>i</sub> site.<ref name=":1" /> The total reaction is shown:

<math chem>2\ \underbrace{\mathrm{cytochrome\ c}}_{\mathrm{oxidized}}\ +\ \ce{UQH_2}\ +\ 2\underbrace{\ce{H^+}}_{\mathrm{matrix}} \longrightarrow 2\ \underbrace{\mathrm{cytochrome\ c}}_{\mathrm{reduced}}\ +\ \ce{UQ}\ +\ 4\underbrace{\ce{H^+}}_{\mathrm{IMS}}</math><ref name=":1" />


Complex IV (CIV) catalyzes the transfer of two electrons from the cytochrome c reduced by CIII to one half of a full oxygen. Utilizing one full oxygen in oxidative phosphorylation requires the transfer of four electrons. The oxygen will then consume four protons from the matrix to form water while another four protons are pumped into the IMS.<ref name=":1" /> The total reaction is shown:
[[Complex III]] (CIII) catalyzes the [[Q cycle|Q-cycle]]. The first step involving the transfer of two electrons from the UQH<sub>2</sub> reduced by CI to two molecules of oxidized [[cytochrome c]] at the Q<sub>o</sub> site. In the second step, two more electrons reduce UQ to UQH<sub>2</sub> at the Q<sub>i</sub> site. The total reaction is:<ref name="PumpStructures" />
<math chem display=block>2\underbrace{\text{cytochrome c}}_{\text{oxidized}}+\ce{UQH_2}+2\underbrace{\ce{H^+}}_{\text{matrix}}\longrightarrow2\underbrace{\text{cytochrome c}}_{\text{reduced}}+\ce{UQ}+4\underbrace{\ce{H^+}}_{\text{IMS}}</math>


Complex IV (CIV) catalyzes the transfer of two electrons from the cytochrome c reduced by CIII to one half of a full oxygen. Utilizing one full oxygen in oxidative phosphorylation requires the transfer of four electrons. The oxygen will then consume four protons from the matrix to form water while another four protons are pumped into the IMS, to give a total reaction<ref name="PumpStructures" />
<math>2\ cytochrome\ c\ (reduced)\ +\ 4H^+\ (matrix)\ +\ 1/2\ O_2 \longrightarrow 2\ cytochrome\ c\ (oxidized)\ +\ 2H^+\ (IMS)\ +\ H_2O</math><ref name=":1" />


== <math chem="" display="block">2\text{cytochrome c}(\text{reduced})+4\ce{H+}(\text{matrix})+\frac{1}{2}\ce{O2}\longrightarrow2\text{cytochrome c}(\text{oxidized})+2\ce{H+}(\text{IMS})+\ce{H2O}</math>See also ==
==See also==
{{Div col|colwidth=20em}}
{{Div col|colwidth=20em}}
*[[Concentration cell]]
*[[Concentration cell]]
Line 186: Line 129:
{{DEFAULTSORT:Electrochemical Gradient}}
{{DEFAULTSORT:Electrochemical Gradient}}
[[Category:Cellular respiration]]
[[Category:Cellular respiration]]
[[Category:Electrochemistry]]
[[Category:Electrochemical concepts]]
[[Category:Electrophysiology]]
[[Category:Electrophysiology]]
[[Category:Membrane biology]]
[[Category:Membrane biology]]

Revision as of 14:53, 18 December 2023

Diagram of ion concentrations and charge across a semi-permeable cellular membrane.

An electrochemical gradient is a gradient of electrochemical potential, usually for an ion that can move across a membrane. The gradient consists of two parts:

  • The chemical gradient, or difference in solute concentration across a membrane.
  • The electrical gradient, or difference in charge across a membrane.

When there are unequal concentrations of an ion across a permeable membrane, the ion will move across the membrane from the area of higher concentration to the area of lower concentration through simple diffusion. Ions also carry an electric charge that forms an electric potential across a membrane. If there is an unequal distribution of charges across the membrane, then the difference in electric potential generates a force that drives ion diffusion until the charges are balanced on both sides of the membrane.

Electrochemical gradients are essential to the operation of batteries and other electrochemical cells, photosynthesis and cellular respiration, and certain other biological processes.

Overview

Electrochemical energy is one of the many interchangeable forms of potential energy through which energy may be conserved. It appears in electroanalytical chemistry and has industrial applications such as batteries and fuel cells. In biology, electrochemical gradients allow cells to control the direction ions move across membranes. In mitochondria and chloroplasts, proton gradients generate a chemiosmotic potential used to synthesize ATP,[1] and the sodium-potassium gradient helps neural synapses quickly transmit information.[citation needed]

An electrochemical gradient has two components: a differential concentration of electric charge across a membrane and a differential concentration of chemical species across that same membrane. In the former effect, the concentrated charge attracts charges of the opposite sign; in the latter, the concentrated species tends to diffuse across the membrane to an equalize concentrations. The combination of these two phenomena determines the thermodynamically-preferred direction for an ion's movement across the membrane.[2]: 403 [3]

The combined effect can be quantified as a gradient in the thermodynamic electrochemical potential:[citation needed] with

  • μi the chemical potential of the ion species i
  • zi the charge per ion of the species i
  • F, Faraday constant (the electrochemical potential is implicitly measured on a per-mole basis)
  • φ, the local electric potential.

Sometimes, the term "electrochemical potential" is abused to describe the electric potential generated by an ionic concentration gradient; that is, φ.

An electrochemical gradient is analogous to the water pressure across a hydroelectric dam. Routes unblocked by the membrane (e.g. membrane transport protein or electrodes) correspond to turbines that convert the water's potential energy to other forms of physical or chemical energy, and the ions that pass through the membrane correspond to water traveling into the lower river.[tone] Conversely, energy can be used to pump water up into the lake above the dam, and chemical energy can be used to create electrochemical gradients.[4][5]

Chemistry

The term typically applies in electrochemistry, when electrical energy in the form of an applied voltage is used to modulate the thermodynamic favorability of a chemical reaction. In a battery, an electrochemical potential arising from the movement of ions balances the reaction energy of the electrodes. The maximum voltage that a battery reaction can produce is sometimes called the standard electrochemical potential of that reaction.

Biological context

The generation of a transmembrane electrical potential through ion movement across a cell membrane drives biological processes like nerve conduction, muscle contraction, hormone secretion, and sensation. By convention, physiological voltages are measured relative to the extracellular region; a typical animal cell has an internal electrical potential of (−70)–(−50) mV.[2]: 464 

An electrochemical gradient is essential to mitochondrial oxidative phosphorylation. The final step of cellular respiration is the electron transport chain, composed of four complexes embedded in the inner mitochondrial membrane. Complexes I, III, and IV pump protons from the matrix to the intermembrane space (IMS); for every electron pair entering the chain, ten protons translocate into the IMS. The result is an electric potential of more than 200 mV. The resulting flux of protons back into the matrix powers the efforts of ATP synthase to combine inorganic phosphate and ADP.[6][2]: 743–745 

Similar to the electron transport chain, the light-dependent reactions of photosynthesis pump protons into the thylakoid lumen of chloroplasts to drive the synthesis of ATP. The proton gradient can be generated through either noncyclic or cyclic photophosphorylation. Of the proteins that participate in noncyclic photophosphorylation, photosystem II (PSII), plastiquinone, and cytochrome b6f complex directly contribute to generating the proton gradient. For each four photons absorbed by PSII, eight protons are pumped into the lumen.[2]: 769–770 

Several other transporters and ion channels play a role in generating a proton electrochemical gradient. One is TPK3, a potassium channel that is activated by Ca2+ and conducts K+ from the thylakoid lumen to the stroma, which helps establish the electric field. On the other hand, the electro-neutral K+ efflux antiporter (KEA3) transports K+ into the thylakoid lumen and H+ into the stroma, which helps establish the pH gradient.[7]

Ion gradients

Diagram of the Na+-K+-ATPase.

Since the ions are charged, they cannot pass through cellular membranes via simple diffusion. Two different mechanisms can transport the ions across the membrane: active or passive transport.[citation needed]

An example of active transport of ions is the Na+-K+-ATPase (NKA). NKA is powered by the hydrolysis of ATP into ADP and an inorganic phosphate; for every molecule of ATP hydrolized, three Na+ are transported outside and two K+ are transported inside the cell. This makes the inside of the cell more negative than the outside and more specifically generates a membrane potential Vmembrane of about −60 mV.[5]

An example of passive transport is ion fluxes through Na+, K+, Ca2+, and Cl channels. Unlike active transport, passive transport is powered by the arithmetic sum of osmosis (a concentration gradient) and an electric field (the transmembrane potential). Formally, the molar Gibbs free energy change associated with successful transport is[citation needed] where R represents the gas constant, T represents absolute temperature, z is the charge per ion, and F represents the Faraday constant.[2]: 464–465 

In the example of Na+, both terms tend to support transport: the negative electric potential inside the cell attracts the positive ion and since Na+ is concentrated outside the cell, osmosis supports diffusion through the Na+ channel into the cell. In the case of K+, the effect of osmosis is reversed: although external ions are attracted by the negative intracellular potential, entropy seeks to diffuse the ions already concentrated inside the cell. The converse phenomenon (osmosis supports transport, electric potential opposes it) can be achieved for Na+ in cells with abnormal transmembrane potentials: at +70 mV, the Na+ influx halts; at higher potentials, it becomes an efflux.[citation needed]

Common cellular ion concentrations (millimolar)[8][9][10][11]
Ion Mammal Squid axon S. cerevisiae E. coli Sea water
Cell Blood Cell Blood
K+ 100 - 140 4-5 400 10 - 20 300 30 - 300 10
Na+ 5-15 145 50 440 30 10 500
Mg2+ 10 [a]
0.5 - 0.8 [b]
1 - 1.5 50 30 - 100 [a]
0.01 - 1 [b]
50
Ca2+ 10−4 2.2 - 2.6 [c]
1.3 - 1.5 [d]
10−4 - 3×10−4 10 2 3 [a]
10−4 [b]
10
Cl 4 110 40 - 150 560 10 - 200 [e] 500
X (negatively charged proteins) 138 9 300 - 400 5-10
HCO3 12 29
pH 7.1 - 7.3[12] 7.35 to 7.45 [12] (normal arterial blood pH)
6.9 - 7.8 [12] (overall range)
7.2 - 7.8[13] 8.1 - 8.2[14]
  1. ^ a b c Bound
  2. ^ a b c Free
  3. ^ Total
  4. ^ Ionised
  5. ^ Medium dependent

Proton gradients

Proton gradients in particular are important in many types of cells as a form of energy storage. The gradient is usually used to drive ATP synthase, flagellar rotation, or metabolite transport.[15] This section will focus on three processes that help establish proton gradients in their respective cells: bacteriorhodopsin and noncyclic photophosphorylation and oxidative phosphorylation.[citation needed]

Bacteriorhodopsin

Diagram of the conformational shift in retinal that initiates proton pumping in bacteriorhodopsin.

The way bacteriorhodopsin generates a proton gradient in Archaea is through a proton pump. The proton pump relies on proton carriers to drive protons from the side of the membrane with a low H+ concentration to the side of the membrane with a high H+ concentration. In bacteriorhodopsin, the proton pump is activated by absorption of photons of 568nm wavelength, which leads to isomerization of the Schiff base (SB) in retinal forming the K state. This moves SB away from Asp85 and Asp212, causing H+ transfer from the SB to Asp85 forming the M1 state. The protein then shifts to the M2 state by separating Glu204 from Glu194 which releases a proton from Glu204 into the external medium. The SB is reprotonated by Asp96 which forms the N state. It is important that the second proton comes from Asp96 since its deprotonated state is unstable and rapidly reprotonated with a proton from the cytosol. The protonation of Asp85 and Asp96 causes re-isomerization of the SB, forming the O state. Finally, bacteriorhodopsin returns to its resting state when Asp85 releases its proton to Glu204.[15][16]

Photophosphorylation

Simplified diagram of photophosphorylation.

PSII also relies on light to drive the formation of proton gradients in chloroplasts, however, PSII utilizes vectorial redox chemistry to achieve this goal. Rather than physically transporting protons through the protein, reactions requiring the binding of protons will occur on the extracellular side while reactions requiring the release of protons will occur on the intracellular side. Absorption of photons of 680nm wavelength is used to excite two electrons in P680 to a higher energy level. These higher energy electrons are transferred to protein-bound plastoquinone (PQA) and then to unbound plastoquinone (PQB). This reduces plastoquinone (PQ) to plastoquinol (PQH2) which is released from PSII after gaining two protons from the stroma. The electrons in P680 are replenished by oxidizing water through the oxygen-evolving complex (OEC). This results in release of O2 and H+ into the lumen, for a total reaction of[15]

After being released from PSII, PQH2 travels to the cytochrome b6f complex, which then transfers two electrons from PQH2 to plastocyanin in two separate reactions. The process that occurs is similar to the Q-cycle in Complex III of the electron transport chain. In the first reaction, PQH2 binds to the complex on the lumen side and one electron is transferred to the iron-sulfur center which then transfers it to cytochrome f which then transfers it to plastocyanin. The second electron is transferred to heme bL which then transfers it to heme bH which then transfers it to PQ. In the second reaction, a second PQH2 gets oxidized, adding an electron to another plastocyanin and PQ. Both reactions together transfer four protons into the lumen.[2]: 782–783 [17]

Oxidative phosphorylation

Detailed diagram of the electron transport chain in mitochondria.

In the electron transport chain, complex I (CI) catalyzes the reduction of ubiquinone (UQ) to ubiquinol (UQH2) by the transfer of two electrons from reduced nicotinamide adenine dinucleotide (NADH) which translocates four protons from the mitochondrial matrix to the IMS:[18]

Complex III (CIII) catalyzes the Q-cycle. The first step involving the transfer of two electrons from the UQH2 reduced by CI to two molecules of oxidized cytochrome c at the Qo site. In the second step, two more electrons reduce UQ to UQH2 at the Qi site. The total reaction is:[18]

Complex IV (CIV) catalyzes the transfer of two electrons from the cytochrome c reduced by CIII to one half of a full oxygen. Utilizing one full oxygen in oxidative phosphorylation requires the transfer of four electrons. The oxygen will then consume four protons from the matrix to form water while another four protons are pumped into the IMS, to give a total reaction[18]

See also

References

  1. ^ Nath, Sunil; Villadsen, John (2015-03-01). "Oxidative phosphorylation revisited". Biotechnology and Bioengineering. 112 (3): 429–437. doi:10.1002/bit.25492. ISSN 1097-0290. PMID 25384602. S2CID 2598635.
  2. ^ a b c d e f Nelson, David; Cox, Michael (2013). Lehninger Principles of Biochemistry. New York: W.H. Freeman. ISBN 978-1-4292-3414-6.
  3. ^ Yang, Huanghe; Zhang, Guohui; Cui, Jianmin (2015-01-01). "BK channels: multiple sensors, one activation gate". Frontiers in Physiology. 6: 29. doi:10.3389/fphys.2015.00029. PMC 4319557. PMID 25705194.
  4. ^ Shattock, Michael J.; Ottolia, Michela; Bers, Donald M.; Blaustein, Mordecai P.; Boguslavskyi, Andrii; Bossuyt, Julie; Bridge, John H. B.; Chen-Izu, Ye; Clancy, Colleen E. (2015-03-15). "Na+/Ca2+ exchange and Na+/K+-ATPase in the heart". The Journal of Physiology. 593 (6): 1361–1382. doi:10.1113/jphysiol.2014.282319. ISSN 1469-7793. PMC 4376416. PMID 25772291.
  5. ^ a b Aperia, Anita; Akkuratov, Evgeny E.; Fontana, Jacopo Maria; Brismar, Hjalmar (2016-04-01). "Na+-K+-ATPase, a new class of plasma membrane receptors". American Journal of Physiology. Cell Physiology. 310 (7): C491–C495. doi:10.1152/ajpcell.00359.2015. ISSN 0363-6143. PMID 26791490.
  6. ^ Poburko, Damon; Demaurex, Nicolas (2012-04-24). "Regulation of the mitochondrial proton gradient by cytosolic Ca2+ signals" (PDF). Pflügers Archiv: European Journal of Physiology. 464 (1): 19–26. doi:10.1007/s00424-012-1106-y. ISSN 0031-6768. PMID 22526460. S2CID 18133149.
  7. ^ Höhner, Ricarda; Aboukila, Ali; Kunz, Hans-Henning; Venema, Kees (2016-01-01). "Proton Gradients and Proton-Dependent Transport Processes in the Chloroplast". Frontiers in Plant Science. 7: 218. doi:10.3389/fpls.2016.00218. PMC 4770017. PMID 26973667.
  8. ^ Philips, Ron Milo & Ron. "» What are the concentrations of different ions in cells?". Retrieved 2019-06-07.
  9. ^ Lodish, Harvey; Berk, Arnold; Zipursky, S. Lawrence; Matsudaira, Paul; Baltimore, David; Darnell, James (2000). "Table 15-1, Typical Ion Concentrations in Invertebrates and Vertebrates". www.ncbi.nlm.nih.gov. Retrieved 2019-06-07.
  10. ^ "The following table gives an idea of the intra and extra cellular ion concentrations in a squid axon and a mammalian cell". www.chm.bris.ac.uk. Retrieved 2019-06-07.
  11. ^ Diem K, Lenter C. Scientific Tables. Vol. 565 (Seventh ed.). Basel: Ciba-Geigy Limited. pp. 653–654. ISBN 978-3-9801244-0-9.
  12. ^ a b c Spitzer, Kenneth W.; Vaughan-Jones, Richard D. (2003), Karmazyn, Morris; Avkiran, Metin; Fliegel, Larry (eds.), "Regulation of Intracellular pH in Mammalian Cells", The Sodium-Hydrogen Exchanger: From Molecule to its Role in Disease, Springer US, pp. 1–15, doi:10.1007/978-1-4615-0427-6_1, ISBN 9781461504276
  13. ^ Slonczewski, Joan L.; Wilks, Jessica C. (2007-08-01). "pH of the Cytoplasm and Periplasm of Escherichia coli: Rapid Measurement by Green Fluorescent Protein Fluorimetry". Journal of Bacteriology. 189 (15): 5601–5607. doi:10.1128/JB.00615-07. ISSN 0021-9193. PMC 1951819. PMID 17545292.
  14. ^ Brewer, Peter G. (September 1, 2008). "Rising Acidity in the Ocean: The Other CO2 Problem". doi:10.1038/scientificamericanearth0908-22. {{cite journal}}: Cite journal requires |journal= (help)
  15. ^ a b c Gunner, M. R.; Amin, Muhamed; Zhu, Xuyu; Lu, Jianxun (2013-08-01). "Molecular mechanisms for generating transmembrane proton gradients". Biochimica et Biophysica Acta (BBA) - Bioenergetics. Metals in Bioenergetics and Biomimetics Systems. 1827 (8–9): 892–913. doi:10.1016/j.bbabio.2013.03.001. PMC 3714358. PMID 23507617.
  16. ^ Wickstrand, Cecilia; Dods, Robert; Royant, Antoine; Neutze, Richard (2015-03-01). "Bacteriorhodopsin: Would the real structural intermediates please stand up?". Biochimica et Biophysica Acta (BBA) - General Subjects. Structural biochemistry and biophysics of membrane proteins. 1850 (3): 536–553. doi:10.1016/j.bbagen.2014.05.021. PMID 24918316.
  17. ^ Schöttler, Mark Aurel; Tóth, Szilvia Z.; Boulouis, Alix; Kahlau, Sabine (2015-05-01). "Photosynthetic complex stoichiometry dynamics in higher plants: biogenesis, function, and turnover of ATP synthase and the cytochrome b 6 f complex". Journal of Experimental Botany. 66 (9): 2373–2400. doi:10.1093/jxb/eru495. ISSN 0022-0957. PMID 25540437.
  18. ^ a b c Sun, Fei; Zhou, Qiangjun; Pang, Xiaoyun; Xu, Yingzhi; Rao, Zihe (2013-08-01). "Revealing various coupling of electron transfer and proton pumping in mitochondrial respiratory chain". Current Opinion in Structural Biology. 23 (4): 526–538. doi:10.1016/j.sbi.2013.06.013. PMID 23867107.