Home  

Random  

Nearby  



Log in  



Settings  



Donate  



About Wikipedia  

Disclaimers  



Wikipedia





Ideal (ring theory)





Article  

Talk  



Language  

Watch  

Edit  





Inmathematics, 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.

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 idealinorder 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

edit

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 and motivation

edit

For an arbitrary ring  , let   be its additive group. A subset I is called a left idealof  if it is an additive subgroup 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 subgroupof ,
  2. For every   and every  , the product   is in  .

Aright 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. In the language of modules, the definitions mean that a left (resp. right, two-sided) ideal of   is an  -submoduleof  when   is viewed as a left (resp. right, bi-)  -module. When   is a commutative ring, the definitions of left, right, and two-sided ideal coincide, and the term ideal is used alone.

To understand the concept of an ideal, consider how ideals arise in the construction of rings of "elements modulo". For concreteness, let us look at the ring   of integers modulo   given an integer   (  is a commutative ring). The key observation here is that we obtain   by taking the integer line   and wrapping it around itself so that various integers get identified. In doing so, we must satisfy two requirements:

  1.   must be identified with 0 since   is congruent to 0 modulo  .
  2. the resulting structure must again be a ring.

The second requirement forces us to make additional identifications (i.e., it determines the precise way in which we must wrap   around itself). The notion of an ideal arises when we ask the question:

What is the exact set of integers that we are forced to identify with 0?

The answer is, unsurprisingly, the set   of all integers congruent to 0 modulo  . That is, we must wrap   around itself infinitely many times so that the integers   will all align with 0. If we look at what properties this set must satisfy in order to ensure that   is a ring, then we arrive at the definition of an ideal. Indeed, one can directly verify that   is an ideal of  .

Remark. Identifications with elements other than 0 also need to be made. For example, the elements in   must be identified with 1, the elements in   must be identified with 2, and so on. Those, however, are uniquely determined by   since   is an additive group.

We can make a similar construction in any commutative ring  : start with an arbitrary  , and then identify with 0 all elements of the ideal  . It turns out that the ideal   is the smallest ideal that contains  , called the ideal generatedby . More generally, we can start with an arbitrary subset  , and then identify with 0 all the elements in the ideal generated by  : the smallest ideal   such that  . The ring that we obtain after the identification depends only on the ideal   and not on the set   that we started with. That is, if  , then the resulting rings will be the same.

Therefore, an ideal   of a commutative ring   captures canonically the information needed to obtain the ring of elements of   modulo a given subset  . The elements of  , by definition, are those that are congruent to zero, that is, identified with zero in the resulting ring. The resulting ring is called the quotientof by  and is denoted  . Intuitively, the definition of an ideal postulates two natural conditions necessary for   to contain all elements designated as "zeros" by  :

  1.   is an additive subgroup of  : the zero 0 of   is a "zero"  , and if   and   are "zeros", then   is a "zero" too.
  2. Any   multiplied by a "zero"   is a "zero"  .

It turns out that the above conditions are also sufficient for   to contain all the necessary "zeros": no other elements have to be designated as "zero" in order to form  . (In fact, no other elements should be designated as "zero" if we want to make the fewest identifications.)

Remark. The above construction still works using two-sided ideals even if   is not necessarily commutative.

Examples and properties

edit

(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

edit

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

edit

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 ain  and bin .

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:  .[11])

An integral domain is called a Dedekind domain if for each pair of ideals  , there is an ideal   such that  .[12] 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

edit

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.[13][14][15]

Radical of a ring

edit

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 idealofR is the annihilator of a (nonzero) simple R-module. The Jacobson radical  ofR 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 nilradicalofR. As it turns out,   is also the set of nilpotent elementsofR.

IfR 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

edit

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 inB 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 toA.

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 fissurjective and   then:

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

The following is sometimes useful:[16] 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  ofB, disjoint from  , such that   is a maximal ideal containing  . One then checks that   lies over  . The converse is obvious.)

Generalizations

edit

Ideals can be generalized to any monoid object  , where   is the object where the monoid structure has been forgotten. A left idealof  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 subobjectof 
  2. For every   and every  , the product   is in  .

Aright 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-moduleof . 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

edit

Notes

edit
  1. ^ Some authors call the zero and unit ideals of a ring R the trivial idealsofR.
  • ^ IfR 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 xinX and every n in the natural numbers. When R has a unit, this extra requirement becomes superfluous.
  • References

    edit
    1. ^ a b John Stillwell (2010). Mathematics and its history. p. 439.
  • ^ Harold M. Edwards (1977). Fermat's last theorem. A genetic introduction to algebraic number theory. p. 76.
  • ^ Everest G., Ward T. (2005). An introduction to number theory. p. 83.
  • ^ a b c Dummit & Foote (2004), p. 243.
  • ^ Lang 2005, Section III.2
  • ^ Dummit & Foote (2004), p. 244.
  • ^ Because simple commutative rings are fields. See Lam (2001). A First Course in Noncommutative Rings. p. 39.
  • ^ Dummit & Foote (2004), p. 255.
  • ^ Dummit & Foote (2004), p. 251.
  • ^ Matsumura, Hideyuki (1987). Commutative Ring Theory. Cambridge: Cambridge University Press. p. 132. ISBN 9781139171762.
  • ^ Eisenbud 1995, Exercise A 3.17
  • ^ Milnor (1971), p. 9.
  • ^ "ideals". www.math.uiuc.edu. Archived from the original on 2017-01-16. Retrieved 2017-01-14.
  • ^ "sums, products, and powers of ideals". www.math.uiuc.edu. Archived from the original on 2017-01-16. Retrieved 2017-01-14.
  • ^ "intersection of ideals". www.math.uiuc.edu. Archived from the original on 2017-01-16. Retrieved 2017-01-14.
  • ^ Atiyah & Macdonald (1969), Proposition 3.16.
  • Dummit, David Steven; Foote, Richard Martin (2004). Abstract algebra (Third ed.). Hoboken, NJ: John Wiley & Sons, Inc. ISBN 9780471433347.
  • Eisenbud, David (1995), Commutative Algebra with a View toward Algebraic Geometry, Graduate Texts in Mathematics, vol. 150, Berlin, New York: Springer-Verlag, doi:10.1007/978-1-4612-5350-1, ISBN 978-0-387-94268-1, MR 1322960
  • Lang, Serge (2005). Undergraduate Algebra (Third ed.). Springer-Verlag. ISBN 978-0-387-22025-3.
  • Hazewinkel, Michiel; Gubareni, Nadiya; Gubareni, Nadezhda Mikhaĭlovna; Kirichenko, Vladimir V. (2004). Algebras, rings and modules. Vol. 1. Springer. ISBN 1-4020-2690-0.
  • Milnor, John Willard (1971). Introduction to algebraic K-theory. Annals of Mathematics Studies. Vol. 72. Princeton, NJ: Princeton University Press. ISBN 9780691081014. MR 0349811. Zbl 0237.18005.
  • edit

    Retrieved from "https://en.wikipedia.org/w/index.php?title=Ideal_(ring_theory)&oldid=1218915733"
     



    Last edited on 14 April 2024, at 17:06  





    Languages

     


    العربية
    Български
    Català
    Чӑвашла
    Čeština
    Dansk
    Deutsch
    Eesti
    Ελληνικά
    Español
    Esperanto
    فارسی
    Français

    Bahasa Indonesia
    Interlingua
    Italiano
    עברית
    Қазақша
    Lombard
    Magyar
    Nederlands

    Norsk bokmål
    Norsk nynorsk
    Polski
    Português
    Română
    Русский
    Slovenčina
    Slovenščina
    Српски / srpski
    Suomi
    Svenska
    ி
    Українська
    Tiếng Vit


     

    Wikipedia


    This page was last edited on 14 April 2024, at 17:06 (UTC).

    Content is available under CC BY-SA 4.0 unless otherwise noted.



    Privacy policy

    About Wikipedia

    Disclaimers

    Contact Wikipedia

    Code of Conduct

    Developers

    Statistics

    Cookie statement

    Terms of Use

    Desktop