Ideal (ring theory)

Last updated

In mathematics, and more specifically in ring theory, an ideal of a ring is a special subset of its elements. Ideals generalize certain subsets of the integers, such as the even numbers or the multiples of 3. Addition and subtraction of even numbers preserves evenness, and multiplying an even number by any integer (even or odd) results in an even number; these closure and absorption properties are the defining properties of an ideal. An ideal can be used to construct a quotient ring in a way similar to how, in group theory, a normal subgroup can be used to construct a quotient group.

Contents

Among the integers, the ideals correspond one-for-one with the non-negative integers: in this ring, every ideal is a principal ideal consisting of the multiples of a single non-negative number. However, in other rings, the ideals may not correspond directly to the ring elements, and certain properties of integers, when generalized to rings, attach more naturally to the ideals than to the elements of the ring. For instance, the prime ideals of a ring are analogous to prime numbers, and the Chinese remainder theorem can be generalized to ideals. There is a version of unique prime factorization for the ideals of a Dedekind domain (a type of ring important in number theory).

The related, but distinct, concept of an ideal in order theory is derived from the notion of ideal in ring theory. A fractional ideal is a generalization of an ideal, and the usual ideals are sometimes called integral ideals for clarity.

History

Ernst Kummer invented the concept of ideal numbers to serve as the "missing" factors in number rings in which unique factorization fails; here the word "ideal" is in the sense of existing in imagination only, in analogy with "ideal" objects in geometry such as points at infinity. [1] In 1876, Richard Dedekind replaced Kummer's undefined concept by concrete sets of numbers, sets that he called ideals, in the third edition of Dirichlet's book Vorlesungen über Zahlentheorie , to which Dedekind had added many supplements. [1] [2] [3] Later the notion was extended beyond number rings to the setting of polynomial rings and other commutative rings by David Hilbert and especially Emmy Noether.

Definitions

Given a ring R, a left ideal is a subset I of R that is a subgroup of the additive group of that "absorbs multiplication from the left by elements of "; that is, is a left ideal if it satisfies the following two conditions:

  1. is a subgroup of ,
  2. For every and every , the product is in . [4]

In other words, a left ideal is a left submodule of R, considered as a left module over itself. [5]

A right ideal is defined similarly, with the condition replaced by . A two-sided ideal is a left ideal that is also a right ideal.

If the ring is commutative, the three definitions are the same, and one talks simply of an ideal. In the non-commutative case, "ideal" is often used instead of "two-sided ideal".

If I is a left, right or two-sided ideal, the relation if and only if

is an equivalence relation on R, and the set of equivalence classes forms a left, right or bi module denoted and called the quotient of R by I. [6] (It is an instance of a congruence relation and is a generalization of modular arithmetic.)

If the ideal I is two-sided, is a ring, [7] and the function

that associates to each element of R its equivalence class is a surjective ring homomorphism that has the ideal as its kernel. [8] Conversely, the kernel of a ring homomorphism is a two-sided ideal. Therefore, the two-sided ideals are exactly the kernels of ring homomorphisms.

Note on convention

By convention, a ring has the multiplicative identity. But some authors do not require a ring to have the multiplicative identity; i.e., for them, a ring is a rng. For a rng R, a left idealI is a subrng with the additional property that is in I for every and every . (Right and two-sided ideals are defined similarly.) For a ring, an ideal I (say a left ideal) is rarely a subring; since a subring shares the same multiplicative identity with the ambient ring R, if I were a subring, for every , we have i.e., .

The notion of an ideal does not involve associativity; thus, an ideal is also defined for non-associative rings (often without the multiplicative identity) such as a Lie algebra.

Examples and properties

(For the sake of brevity, some results are stated only for left ideals but are usually also true for right ideals with appropriate notation changes.)

(since such a span is the smallest left ideal containing X.) [note 2] A right (resp. two-sided) ideal generated by X is defined in the similar way. For "two-sided", one has to use linear combinations from both sides; i.e.,

Types of ideals

To simplify the description all rings are assumed to be commutative. The non-commutative case is discussed in detail in the respective articles.

Ideals are important because they appear as kernels of ring homomorphisms and allow one to define factor rings. Different types of ideals are studied because they can be used to construct different types of factor rings.

Two other important terms using "ideal" are not always ideals of their ring. See their respective articles for details:

Ideal operations

The sum and product of ideals are defined as follows. For and , left (resp. right) ideals of a ring R, their sum is

,

which is a left (resp. right) ideal, and, if are two-sided,

i.e. the product is the ideal generated by all products of the form ab with a in and b in .

Note is the smallest left (resp. right) ideal containing both and (or the union ), while the product is contained in the intersection of and .

The distributive law holds for two-sided ideals ,

If a product is replaced by an intersection, a partial distributive law holds:

where the equality holds if contains or .

Remark: The sum and the intersection of ideals is again an ideal; with these two operations as join and meet, the set of all ideals of a given ring forms a complete modular lattice. The lattice is not, in general, a distributive lattice. The three operations of intersection, sum (or join), and product make the set of ideals of a commutative ring into a quantale.

If are ideals of a commutative ring R, then in the following two cases (at least)

(More generally, the difference between a product and an intersection of ideals is measured by the Tor functor: . [17] )

An integral domain is called a Dedekind domain if for each pair of ideals , there is an ideal such that . [18] It can then be shown that every nonzero ideal of a Dedekind domain can be uniquely written as a product of maximal ideals, a generalization of the fundamental theorem of arithmetic.

Examples of ideal operations

In we have

since is the set of integers that are divisible by both and .

Let and let . Then,

In the first computation, we see the general pattern for taking the sum of two finitely generated ideals, it is the ideal generated by the union of their generators. In the last three we observe that products and intersections agree whenever the two ideals intersect in the zero ideal. These computations can be checked using Macaulay2. [19] [20] [21]

Radical of a ring

Ideals appear naturally in the study of modules, especially in the form of a radical.

For simplicity, we work with commutative rings but, with some changes, the results are also true for non-commutative rings.

Let R be a commutative ring. By definition, a primitive ideal of R is the annihilator of a (nonzero) simple R-module. The Jacobson radical of R is the intersection of all primitive ideals. Equivalently,

Indeed, if is a simple module and x is a nonzero element in M, then and , meaning is a maximal ideal. Conversely, if is a maximal ideal, then is the annihilator of the simple R-module . There is also another characterization (the proof is not hard):

For a not-necessarily-commutative ring, it is a general fact that is a unit element if and only if is (see the link) and so this last characterization shows that the radical can be defined both in terms of left and right primitive ideals.

The following simple but important fact (Nakayama's lemma) is built-in to the definition of a Jacobson radical: if M is a module such that , then M does not admit a maximal submodule, since if there is a maximal submodule , and so , a contradiction. Since a nonzero finitely generated module admits a maximal submodule, in particular, one has:

If and M is finitely generated, then .

A maximal ideal is a prime ideal and so one has

where the intersection on the left is called the nilradical of R. As it turns out, is also the set of nilpotent elements of R.

If R is an Artinian ring, then is nilpotent and . (Proof: first note the DCC implies for some n. If (DCC) is an ideal properly minimal over the latter, then . That is, , a contradiction.)

Extension and contraction of an ideal

Let A and B be two commutative rings, and let f : AB be a ring homomorphism. If is an ideal in A, then need not be an ideal in B (e.g. take f to be the inclusion of the ring of integers Z into the field of rationals Q). The extension of in B is defined to be the ideal in B generated by . Explicitly,

If is an ideal of B, then is always an ideal of A, called the contraction of to A.

Assuming f : AB is a ring homomorphism, is an ideal in A, is an ideal in B, then:

It is false, in general, that being prime (or maximal) in A implies that is prime (or maximal) in B. Many classic examples of this stem from algebraic number theory. For example, embedding . In , the element 2 factors as where (one can show) neither of are units in B. So is not prime in B (and therefore not maximal, as well). Indeed, shows that , , and therefore .

On the other hand, if f is surjective and then:

Remark: Let K be a field extension of L, and let B and A be the rings of integers of K and L, respectively. Then B is an integral extension of A, and we let f be the inclusion map from A to B. The behaviour of a prime ideal of A under extension is one of the central problems of algebraic number theory.

The following is sometimes useful: [22] a prime ideal is a contraction of a prime ideal if and only if . (Proof: Assuming the latter, note intersects , a contradiction. Now, the prime ideals of correspond to those in B that are disjoint from . Hence, there is a prime ideal of B, disjoint from , such that is a maximal ideal containing . One then checks that lies over . The converse is obvious.)

Generalizations

Ideals can be generalized to any monoid object , where is the object where the monoid structure has been forgotten. A left ideal of is a subobject that "absorbs multiplication from the left by elements of "; that is, is a left ideal if it satisfies the following two conditions:

  1. is a subobject of
  2. For every and every , the product is in .

A right ideal is defined with the condition "" replaced by "'". A two-sided ideal is a left ideal that is also a right ideal, and is sometimes simply called an ideal. When is a commutative monoid object respectively, the definitions of left, right, and two-sided ideal coincide, and the term ideal is used alone.

An ideal can also be thought of as a specific type of R-module. If we consider as a left -module (by left multiplication), then a left ideal is really just a left sub-module of . In other words, is a left (right) ideal of if and only if it is a left (right) -module that is a subset of . is a two-sided ideal if it is a sub--bimodule of .

Example: If we let , an ideal of is an abelian group that is a subset of , i.e. for some . So these give all the ideals of .

See also

Notes

  1. Some authors call the zero and unit ideals of a ring R the trivial ideals of R.
  2. If R does not have a unit, then the internal descriptions above must be modified slightly. In addition to the finite sums of products of things in X with things in R, we must allow the addition of n-fold sums of the form x + x + ... + x, and n-fold sums of the form (−x) + (−x) + ... + (−x) for every x in X and every n in the natural numbers. When R has a unit, this extra requirement becomes superfluous.

Related Research Articles

In mathematics, an integral domain is a nonzero commutative ring in which the product of any two nonzero elements is nonzero. Integral domains are generalizations of the ring of integers and provide a natural setting for studying divisibility. In an integral domain, every nonzero element a has the cancellation property, that is, if a ≠ 0, an equality ab = ac implies b = c.

<span class="mw-page-title-main">Prime ideal</span> Ideal in a ring which has properties similar to prime elements

In algebra, a prime ideal is a subset of a ring that shares many important properties of a prime number in the ring of integers. The prime ideals for the integers are the sets that contain all the multiples of a given prime number, together with the zero ideal.

In commutative algebra, the prime spectrum of a commutative ring R is the set of all prime ideals of R, and is usually denoted by ; in algebraic geometry it is simultaneously a topological space equipped with the sheaf of rings .

In mathematics, rings are algebraic structures that generalize fields: multiplication need not be commutative and multiplicative inverses need not exist. Informally, a ring is a set equipped with two binary operations satisfying properties analogous to those of addition and multiplication of integers. Ring elements may be numbers such as integers or complex numbers, but they may also be non-numerical objects such as polynomials, square matrices, functions, and power series.

In commutative algebra, the Krull dimension of a commutative ring R, named after Wolfgang Krull, is the supremum of the lengths of all chains of prime ideals. The Krull dimension need not be finite even for a Noetherian ring. More generally the Krull dimension can be defined for modules over possibly non-commutative rings as the deviation of the poset of submodules.

In mathematics, Hilbert's Nullstellensatz is a theorem that establishes a fundamental relationship between geometry and algebra. This relationship is the basis of algebraic geometry. It relates algebraic sets to ideals in polynomial rings over algebraically closed fields. This relationship was discovered by David Hilbert, who proved the Nullstellensatz in his second major paper on invariant theory in 1893.

In ring theory, a branch of mathematics, the radical of an ideal of a commutative ring is another ideal defined by the property that an element is in the radical if and only if some power of is in . Taking the radical of an ideal is called radicalization. A radical ideal is an ideal that is equal to its radical. The radical of a primary ideal is a prime ideal.

In commutative algebra and algebraic geometry, localization is a formal way to introduce the "denominators" to a given ring or module. That is, it introduces a new ring/module out of an existing ring/module R, so that it consists of fractions such that the denominator s belongs to a given subset S of R. If S is the set of the non-zero elements of an integral domain, then the localization is the field of fractions: this case generalizes the construction of the field of rational numbers from the ring of integers.

In mathematics, the annihilator of a subset S of a module over a ring is the ideal formed by the elements of the ring that give always zero when multiplied by each element of S.

In mathematics, specifically algebraic geometry, a scheme is a structure that enlarges the notion of algebraic variety in several ways, such as taking account of multiplicities and allowing "varieties" defined over any commutative ring.

In algebra, a module homomorphism is a function between modules that preserves the module structures. Explicitly, if M and N are left modules over a ring R, then a function is called an R-module homomorphism or an R-linear map if for any x, y in M and r in R,

In algebra, flat modules include free modules, projective modules, and, over a principal ideal domain, torsion-free modules. Formally, a module M over a ring R is flat if taking the tensor product over R with M preserves exact sequences. A module is faithfully flat if taking the tensor product with a sequence produces an exact sequence if and only if the original sequence is exact.

In abstract algebra, a valuation ring is an integral domain D such that for every non-zero element x of its field of fractions F, at least one of x or x−1 belongs to D.

In mathematics, the Lasker–Noether theorem states that every Noetherian ring is a Lasker ring, which means that every ideal can be decomposed as an intersection, called primary decomposition, of finitely many primary ideals. The theorem was first proven by Emanuel Lasker for the special case of polynomial rings and convergent power series rings, and was proven in its full generality by Emmy Noether.

In mathematics, ideal theory is the theory of ideals in commutative rings. While the notion of an ideal exists also for non-commutative rings, a much more substantial theory exists only for commutative rings

In mathematics, the Noether normalization lemma is a result of commutative algebra, introduced by Emmy Noether in 1926. It states that for any field k, and any finitely generated commutative k-algebraA, there exist elements y1, y2, ..., yd in A that are algebraically independent over k and such that A is a finitely generated module over the polynomial ring S = k [y1, y2, ..., yd]. The integer d is equal to the Krull dimension of the ring A; and if A is an integral domain, d is also the transcendence degree of the field of fractions of A over k.

In algebraic geometry, a morphism of schemes generalizes a morphism of algebraic varieties just as a scheme generalizes an algebraic variety. It is, by definition, a morphism in the category of schemes.

In commutative algebra, an element b of a commutative ring B is said to be integral over a subring A of B if b is a root of some monic polynomial over A.

In commutative algebra, the support of a module M over a commutative ring R is the set of all prime ideals of R such that . It is denoted by . The support is, by definition, a subset of the spectrum of R.

In commutative algebra, an integrally closed domainA is an integral domain whose integral closure in its field of fractions is A itself. Spelled out, this means that if x is an element of the field of fractions of A that is a root of a monic polynomial with coefficients in A, then x is itself an element of A. Many well-studied domains are integrally closed, as shown by the following chain of class inclusions:

References

  1. 1 2 John Stillwell (2010). Mathematics and its history. p. 439.
  2. Harold M. Edwards (1977). Fermat's last theorem. A genetic introduction to algebraic number theory. p. 76.
  3. Everest G., Ward T. (2005). An introduction to number theory. p. 83.
  4. Dummit & Foote 2004 , p. 242
  5. Dummit & Foote 2004 , § 10.1., Examples (1).
  6. Dummit & Foote 2004 , § 10.1., Proposition 3.
  7. Dummit & Foote 2004 , Ch. 7, Proposition 6.
  8. Dummit & Foote 2004 , Ch. 7, Theorem 7.
  9. 1 2 3 4 Dummit & Foote (2004), p. 243.
  10. Lang 2005 , Section III.2
  11. Dummit & Foote (2004), p. 244.
  12. Because simple commutative rings are fields. See Lam (2001). A First Course in Noncommutative Rings. p. 39.
  13. "Zero ideal". Math World. 22 Aug 2024.
  14. Dummit & Foote (2004), p. 255.
  15. Dummit & Foote (2004), p. 251.
  16. Matsumura, Hideyuki (1987). Commutative Ring Theory. Cambridge: Cambridge University Press. p. 132. ISBN   9781139171762.
  17. Eisenbud 1995 , Exercise A 3.17
  18. Milnor (1971), p. 9.
  19. "ideals". www.math.uiuc.edu. Archived from the original on 2017-01-16. Retrieved 2017-01-14.
  20. "sums, products, and powers of ideals". www.math.uiuc.edu. Archived from the original on 2017-01-16. Retrieved 2017-01-14.
  21. "intersection of ideals". www.math.uiuc.edu. Archived from the original on 2017-01-16. Retrieved 2017-01-14.
  22. Atiyah & Macdonald (1969), Proposition 3.16.