You are on page 1of 33

Algebra II

PDF generated using the open source mwlib toolkit. See http://code.pediapress.com/ for more information. PDF generated at: Mon, 02 Sep 2013 07:16:38 UTC

Contents
Articles
Structure theorem for finitely generated modules over a principal ideal domain Torsion (algebra) Zero divisor Smith normal form Finitely-generated module Free module Chinese remainder theorem Bzout's identity 1 5 8 10 14 17 19 27

References
Article Sources and Contributors Image Sources, Licenses and Contributors 29 30

Article Licenses
License 31

Structure theorem for finitely generated modules over a principal ideal domain

Structure theorem for finitely generated modules over a principal ideal domain
In mathematics, in the field of abstract algebra, the structure theorem for finitely generated modules over a principal ideal domain is a generalization of the fundamental theorem of finitely generated abelian groups and roughly states that finitely generated modules can be uniquely decomposed in much the same way that integers have a prime factorization. The result provides a simple framework to understand various canonical form results for square matrices over fields.

Statement
When a vector space over a field F has a finite generating set, then one may extract from it a basis consisting of a finite number n of vectors, and the space is therefore isomorphic to Fn. The corresponding statement with the F generalized to a principal ideal domain R is no longer true, as a finitely generated module over R need not have any basis. However such a module is still isomorphic to a quotient of some module Rn with n finite (to see this it suffices to construct the morphism that sends the elements of the canonical basis Rn to the generators of the module, and take the quotient by its kernel.) By changing the choice of generating set, one can in fact describe the module as the quotient of some Rn by a particularly simple submodule, and this is the structure theorem. The structure theorem for finitely generated modules over a principal ideal domain usually appears in the following two forms.

Invariant factor decomposition


Every finitely generated module M over a principal ideal domain R is isomorphic to a unique one of the form

where

and

. The order of the nonzero

ideals is invariant, and the number of

is invariant. The nonzero elements, together with the number of

which are zero, form a complete set of invariants for the

module. Explicitly, this means that any two modules sharing the same set of invariants are necessarily isomorphic. The themselves are called invariant factors of M. The ideals are unique. In terms of the elements, this means that the are unique up to multiplication by a factors. These occur at the end

unit. The free part is visible in the part of the decomposition corresponding to the of the sequence of 's, as everything divides zero. Some prefer to write the free part of M separately:

where the visible

are nonzero, and f is the number of

's which are 0.

Structure theorem for finitely generated modules over a principal ideal domain

Primary decomposition
Every finitely generated module M over a principal ideal domain R is isomorphic to one of the form

where The elements The summands

and the .

are primary ideals. The

are unique (up to multiplication by units).

are called the elementary divisors of M. In a PID, primary ideals are powers of primes, and so are indecomposable, so the primary decomposition is a decomposition into indecomposable

modules, and thus every finitely generated module over a PID is a completely decomposable module. Since PID's are Noetherian rings, this can be seen as a manifestation of the Lasker-Noether theorem. As before, it is possible to write the free part (where ) separately and express M as:

where the visible

are nonzero.

Proofs
One proof proceeds as follows: Every finitely generated module over a PID is also finitely presented because a PID is Noetherian, an even stronger condition than coherence. Take a presentation, which is a map (relations to generators), and put it in Smith normal form. This yields the invariant factor decomposition, and the diagonal entries of Smith normal form are the invariant factors. Another outline of a proof: Denote by tM the torsion submodule of M. Then M/tM is a finitely generated torsion free module, and such a module over a commutative PID is a free module of finite rank, so it is isomorphic to for a positive integer n. This free module can be embedded as a submodule F of M, such that the embedding splits (is a right inverse of) the projection map; it suffices to lift each of the generators of F into M. As a consequence . For a prime p in R we can then speak of for each prime p. This is a

submodule of tM, and it turns out that each Np is a direct sum of cyclic modules, and that tM is a direct sum of Np for a finite number of distinct primes p. Putting the previous two steps together, M is decomposed into cyclic modules of the indicated types.

Corollaries
This includes the classification of finite-dimensional vector spaces as a special case, where have no non-trivial ideals, every finitely generated vector space is free. Taking yields the fundamental theorem of finitely generated abelian groups. , the algebra of . Since fields

Let T be a linear operator on a finite-dimensional vector space V over K. Taking

polynomials with coefficients in K evaluated at T, yields structure information about T. V can be viewed as a finitely generated module over . The last invariant factor is the minimal polynomial, and the product of invariant factors is the characteristic polynomial. Combined with a standard matrix form for , this yields various canonical forms: invariant factors + companion matrix yields Frobenius normal form (aka, rational canonical form) primary decomposition + companion matrix yields primary rational canonical form

Structure theorem for finitely generated modules over a principal ideal domain primary decomposition + Jordan blocks yields Jordan canonical form (this latter only holds over an algebraically closed field)

Uniqueness
While the invariants (rank, invariant factors, and elementary divisors) are unique, the isomorphism between M and its canonical form is not unique, and does not even preserve the direct sum decomposition. This follows because there are non-trivial automorphisms of these modules which do not preserve the summands. However, one has a canonical torsion submodule T, and similar canonical submodules corresponding to each (distinct) invariant factor, which yield a canonical sequence:

Compare composition series in JordanHlder theorem. For instance, if change of basis matrix , and is one basis, then is another basis, and the summand,

does not preserve the summand

. However, it does preserve the

as this is the torsion submodule (equivalently here, the 2-torsion elements).

Generalizations
Groups
The JordanHlder theorem is a more general result for finite groups (or modules over an arbitrary ring). In this generality, one obtains a composition series, rather than a direct sum. The KrullSchmidt theorem and related results give conditions under which a module has something like a primary decomposition, a decomposition as a direct sum of indecomposable modules in which the summands are unique up to order.

Primary decomposition
The primary decomposition generalizes to finitely generated modules over commutative Noetherian rings, and this result is called the LaskerNoether theorem.

Indecomposable modules
By contrast, unique decomposition into indecomposable submodules does not generalize as far, and the failure is measured by the ideal class group, which vanishes for PIDs. For rings that are not principal ideal domains, unique decomposition need not even hold for modules over a ring generated by two elements. For the ring R=Z[5], both the module R and its submodule M generated by 2 and 1+5 are indecomposable. While R is not isomorphic to M, RR is isomorphic to MM; thus the images of the M summands give indecomposable submodules L1,L2<RR which give a different decomposition of RR. The failure of uniquely factorizing RR into a direct sum of indecomposable modules is directly related (via the ideal class group) to the failure of the unique factorization of elements of R into irreducible elements ofR.

Structure theorem for finitely generated modules over a principal ideal domain

Non-finitely generated modules


Similarly for modules that are not finitely generated, one cannot expect such a nice decomposition: even the number of factors may vary. There are Z-submodules of Q4 which are simultaneously direct sums of two indecomposable modules and direct sums of three indecomposable modules, showing the analogue of the primary decomposition cannot hold for infinitely generated modules, even over the integers, Z. Another issue that arises with non-finitely generated modules is that there are torsion-free modules which are not free. For instance, consider the ring Z of integers. A classical example of a torsion-free module which is not free is the BaerSpecker group, the group of all sequences of integers under termwise addition. In general, the question of which infinitely generated torsion-free abelian groups are free depends on which large cardinals exist. A consequence is that any structure theorem for infinitely generated modules depends on a choice of set theory axioms and may be invalid under a different choice.

References
Atiyah, Michael Francis; Macdonald, I.G. (1969), Introduction to Commutative Algebra, Westview Press, ISBN978-0-201-40751-8 Dummit, David S.; Foote, Richard M. (2004), Abstract algebra (3rd ed.), New York: Wiley, ISBN978-0-471-43334-7, MR 2286236 (http://www.ams.org/mathscinet-getitem?mr=2286236) Hungerford, Thomas W. (1980), Algebra, New York: Springer, pp.218226, Section IV.6: Modules over a Principal Ideal Domain, ISBN978-0-387-90518-1 Jacobson, Nathan (1985), Basic algebra. I (2 ed.), New York: W. H. Freeman and Company, pp.xviii+499, ISBN0-7167-1480-9, MR 780184 (http://www.ams.org/mathscinet-getitem?mr=780184) Lam, T. Y. (1999), Lectures on modules and rings, Graduate Texts in Mathematics No. 189, Springer-Verlag, ISBN978-0-387-98428-5

Torsion (algebra)

Torsion (algebra)
In abstract algebra, the term torsion refers to elements of finite order in groups and to elements of modules annihilated by regular elements of a ring.

Definition
An element m of a module M over a ring R is called a torsion element of the module if there exists a regular element r of the ring (a non-zero element of the ring that is neither a left nor a right zero divisor) that annihilates m, i.e., r m = 0. In an integral domain (a commutative ring without zero divisors), every non-zero element is regular, so a torsion element of a module over an integral domain is one annihilated by a non-zero element of the integral domain. Some authors use this as the definition of a torsion element but this definition does not work well over more general rings. A module M over a ring R is called a torsion module if all its elements are torsion elements, and torsion-free if zero is the only torsion element. If the ring R is commutative then the set of all torsion elements forms a submodule of M, called the torsion submodule of M, sometimes denoted T(M). If R is not commutative, T(M) may or may not be a submodule. It is shown in (Lam 2007) that R is a right Ore ring if and only if T(M) is a submodule of M for all right R modules. Since right Noetherian domains are Ore, this covers the case when R is a right Noetherian domain (which might not be commutative). More generally, let M be a module over a ring R and S be a multiplicatively closed subset of R. An element m of M is called an S-torsion element if there exists an element s in S such that s annihilates m, i.e., s m = 0. In particular, one can take for S the set of regular elements of the ring R and recover the definition above. An element g of a group G is called a torsion element of the group if it has finite order, i.e., if there is a positive integer m such that gm = e, where e denotes the identity element of the group, and gm denotes the product of m copies of g. A group is called a torsion (or periodic) group if all its elements are torsion elements, and a torsion-free group if the only torsion element is the identity element. Any abelian group may be viewed as a module over the ring Z of integers, and in this case the two notions of torsion coincide.

Examples
1. Let M be a free module over any ring R. Then it follows immediately from the definitions that M is torsion-free (if the ring R is not a domain then torsion is considered with respect to the set S of non-zero divisors of R). In particular, any free abelian group is torsion-free and any vector space over a field K is torsion-free when viewed as the module over K. 2. By contrast with Example 1, any finite group (abelian or not) is periodic and finitely generated. Burnside's problem asks whether, conversely, any finitely generated periodic group must be finite. (The answer is "no" in general, even if the period is fixed.) 3. In the modular group, obtained from the group SL(2,Z) of two by two integer matrices with unit determinant by factoring out its center, any nontrivial torsion element either has order two and is conjugate to the element S or has order three and is conjugate to the element ST. In this case, torsion elements do not form a subgroup, for example, SST=T, which has infinite order. 4. The abelian group Q/Z, consisting of the rational numbers (mod 1), is periodic, i.e. every element has finite order. Analogously, the module K(t)/K[t] over the ring R=K[t] of polynomials in one variable is pure torsion. Both these examples can be generalized as follows: if R is a commutative domain and Q is its field of fractions, then Q/R is a torsion R-module. 5. The torsion subgroup of (R/Z,+) is (Q/Z,+) while the groups (R,+),(Z,+) are torsion-free. The quotient of a torsion-free abelian group by a subgroup is torsion-free exactly when the subgroup is a pure subgroup.

Torsion (algebra) 6. Consider a linear operator L acting on a finite-dimensional vector space V. If we view V as an F[L]-module in the natural way, then (as a result of many things, either simply by finite-dimensionality or as a consequence of the CayleyHamilton theorem), V is a torsion F[L]-module.

Case of a principal ideal domain


Suppose that R is a (commutative) principal ideal domain and M is a finitely-generated R-module. Then the structure theorem for finitely generated modules over a principal ideal domain gives a detailed description of the module M up to isomorphism. In particular, it claims that

where F is a free R-module of finite rank (depending only on M) and T(M) is the torsion submodule of M. As a corollary, any finitely-generated torsion-free module over R is free. This corollary does not hold for more general commutative domains, even for R=K[x,y], the ring of polynomials in two variables. For non-finitely generated modules, the above direct decomposition is not true. The torsion subgroup of an abelian group may not be a direct summand of it.

Torsion and localization


Assume that R is a commutative domain and M is an R-module. Let Q be the quotient field of the ring R. Then one can consider the Q-module

obtained from M by extension of scalars. Since Q is a field, a module over Q is a vector space, possibly, infinite-dimensional. There is a canonical homomorphism of abelian groups from M to MQ, and the kernel of this homomorphism is precisely the torsion submodule T(M). More generally, if S is a multiplicatively closed subset of the ring R, then we may consider localization of the R-module M,

which is a module over the localization RS. There is a canonical map from M to MS, whose kernel is precisely the S-torsion submodule of M. Thus the torsion submodule of M can be interpreted as the set of the elements that 'vanish in the localization'. The same interpretation continues to hold in the non-commutative setting for rings satisfying the Ore condition, or more generally for any right denominator set S and right R-module M.

Torsion in homological algebra


The concept of torsion plays an important role in homological algebra. If M and N are two modules over a commutative ring R (for example, two abelian groups, when R=Z), Tor functors yield a family of R-modules Tori(M,N). The S-torsion of an R-module M is canonically isomorphic to Tor1(M,RS/R). The symbol Tor denoting the functors reflects this relation with the algebraic torsion. This same result holds for non-commutative rings as well as long as the set S is a right denominator set.

Torsion (algebra)

Abelian varieties
The torsion elements of an abelian variety are torsion points or, in an older terminology, division points. On elliptic curves they may be computed in terms of division polynomials.

References
Ernst Kunz, "Introduction to Commutative algebra and algebraic geometry", Birkhauser 1985, ISBN 0-8176-3065-1 Irving Kaplansky, "Infinite abelian groups", University of Michigan, 1954. Michiel Hazewinkel (2001), "Torsion submodule" [1], in Hazewinkel, Michiel, Encyclopedia of Mathematics, Springer, ISBN978-1-55608-010-4 Lam, T. Y. (2007), Exercises in modules and rings, Problem Books in Mathematics, New York: Springer, pp.xviii+412, doi:10.1007/978-0-387-48899-8 [2], ISBN0-387-98850-5, MR2278849 [3]
The 4-torsion subgroup of an elliptic curve over the complex numbers.

References
[1] http:/ / www. encyclopediaofmath. org/ index. php?title=T/ t093330 [2] http:/ / dx. doi. org/ 10. 1007%2F978-0-387-48899-8 [3] http:/ / www. ams. org/ mathscinet-getitem?mr=2278849

Zero divisor

Zero divisor
In abstract algebra, two nonzero elements a and b of a ring are respectively called a left zero divisor and a right zero divisor if a b = 0;[1] this is a partial case of divisibility in rings. An element that is a left or a right zero divisor is simply called a zero divisor.[2] An elementw that is both a left and a right zero divisor[3] is called a two-sided zero divisor. If the ring is commutative, then the left and right zero divisors are the same. A non-zero element of a ring that is not a zero divisor is called regular.

Examples
The ring of integers has no zero divisors, but in the ring the number is a zero divisor: as a divisor of , which is a composite number. A nonzero nilpotent element is always a two-sided zero-divisor. Any idempotent element is always a two-sided zero divisor since . An example of a zero divisor in the ring of matrices (over any unital ring except trivial) is the matrix , because for instance Actually, the simplest example of a pair of zero divisor matrices is . A direct product of two or more non-trivial rings always has zero divisors similarly to the just above (the ring of diagonal matrices over a ring is the same as the direct product -matrix example ).

One-sided zero-divisor
Consider the ring of (formal) matrices and then is a left zero divisor iff is even, since with and . Then . If ,

; and it is a right zero

divisor iff is even for similar reasons. If either of is , then it is a two-sided zero-divisor. Here is another example of a ring with an element that is a zero divisor on one side only. Let be the set of all sequences of integers additive group . Take for the ring all additive maps from to , with pointwise addition and composition as the ring operations. (That is, our ring is .) Three examples of elements of this ring are the right shift , the left shift projection map onto the first factor not zero, and the composites zero divisor: the composite , while and are both zero, so to . However, divisor in the ring of additive maps from is a left zero divisor and , and the . All three of these additive maps are is a right zero is not a left is not a right zero divisor and , the endomorphism ring of the

is the identity. Note also that is not in any direction.

is a two-sided zero-divisor since

Zero divisor

Non-examples
The ring of integers modulo a prime number does not have zero divisors and this ring is, in fact, a field, as every non-zero element is a unit. More generally, there are no zero divisors in division rings. A commutative ring with 0 1 and without zero divisors is called an integral domain.

Properties
In the ring of n-by-n matrices over some field, the left and right zero divisors coincide; they are precisely the non-zero singular matrices. In the ring of n-by-n matrices over some integral domain, the zero divisors are precisely the non-zero matrices with determinant zero. Left or right zero divisors can never be units, because if a is invertible and a b = 0, then 0 = a10 = a1a b = b. Every non-trivial idempotent element a in a ring is a zero divisor, since a2 = a implies that a (a 1) = (a 1) a = 0, with nontriviality ensuring that neither factor is 0. Nonzero nilpotent ring elements are also trivially zero divisors. The set of zero divisors is the union of the associated prime ideals of the ring.

Notes
[1] See Hazewinkel et al. (2004), p.2. [2] See Lanski (2005). [3] " is both a left and a right zero divisor" means and , but such and are not necessarily equal.

References
Hazewinkel, Michiel, ed. (2001), "Zero divisor" (http://www.encyclopediaofmath.org/index.php?title=p/ z099230), Encyclopedia of Mathematics, Springer, ISBN978-1-55608-010-4 Michiel Hazewinkel, Nadiya Gubareni, Nadezhda Mikhalovna Gubareni, Vladimir V. Kirichenko. (2004), Algebras, rings and modules, Vol. 1, Springer, ISBN1-4020-2690-0 Charles Lanski (2005), Concepts in Abstract Algebra, American Mathematical Soc., p.342 Weisstein, Eric W., " Zero Divisor (http://mathworld.wolfram.com/ZeroDivisor.html)", MathWorld.

Smith normal form

10

Smith normal form


In mathematics, the Smith normal form is a normal form that can be defined for any matrix (not necessarily square) with entries in a principal ideal domain (PID). The Smith normal form of a matrix is diagonal, and can be obtained from the original matrix by multiplying on the left and right by invertible square matrices. In particular, the integers are a PID, so one can always calculate the Smith normal form of an integer matrix. The Smith normal form is very useful for working with finitely generated modules over a PID, and in particular for deducing the structure of a quotient of a free module.

Definition
Let A be a nonzero mn matrix over a principal ideal domain R. There exist invertible -matrices S, T so that the product S A T is and

and the diagonal elements

satisfy

. This is the Smith normal form of the matrix A. The

elements are unique up to multiplication by a unit and are called the elementary divisors, invariants, or invariant factors. They can be computed (up to multiplication by a unit) as

where A.

(called i-th determinant divisor) equals the greatest common divisor of all

minors of the matrix

Algorithm
Our first goal will be to find invertible square matrices S and T such that the product S A T is diagonal. This is the hardest part of the algorithm and once we have achieved diagonality it becomes relatively easy to put the matrix in Smith normal form. Phrased more abstractly, the goal is to show that, thinking of A as a map from (the free R-module of rank n) to such that (the free R-module of rank m), there are isomorphisms and has the simple form of a diagonal matrix. The matrices S and T can be found

by starting out with identity matrices of the appropriate size, and modifying S each time a row operation is performed on A in the algorithm by the same row operation, and similarly modifying T for each column operation performed. Since row operations are left-multiplications and column operations are right-multiplications, this preserves the invariant where denote current values and A denotes the original matrix; eventually the matrices in this invariant become diagonal. Only invertible row and column operations are performed, which ensures that S and T remain invertible matrices. For a in R \ {0}, write (a) for the number of prime factors of a (these exist and are unique since any PID is also a unique factorization domain). In particular, R is also a Bzout domain, so it is a gcd domain and the gcd of any two elements satisfies a Bzout's identity. To put a matrix into Smith normal form, one can repeatedly apply the following, where t loops from 1 to m.

Smith normal form

11

Step I: Choosing a pivot


Choose jt to be the smallest column index of A with a non-zero entry, starting the search at column index jt-1+1 if t > 1. We wish to have ; if this is the case this step is complete, otherwise there is by assumption some k with .

, and we can exchange rows and k, thereby obtaining Our chosen pivot is now at position (t, jt).

Step II: Improving the pivot


If there is an entry at position (k,jt) such that Bzout property that there exist , in R such that By left-multiplication with an appropriate invertible matrix L, it can be achieved that row t of the matrix product is the sum of times the original row t and times the original row k, that row k of the product is another linear combination of those original rows, and that all other rows are unchanged. Explicitly, if and satisfy the above equation, then for and (which divisions are possible by the definition of ) one has , then, letting , we know by the

so that the matrix

is invertible, with inverse

Now L can be obtained by fitting

into rows and columns t and k of the identity matrix. By construction the

matrix obtained after left-multiplying by L has entry at position (t,jt) (and due to our choice of and it also has an entry 0 at position (k,jt), which is useful though not essential for the algorithm). This new entry divides the entry that was there before, and so in particular ; therefore repeating these steps must eventually terminate. One ends up with a matrix having an entry at position (t,jt) that divides all entries in column jt.

Step III: Eliminating entries


Finally, adding appropriate multiples of row t, it can be achieved that all entries in column jt except for that at position (t,jt) are zero. This can be achieved by left-multiplication with an appropriate matrix. However, to make the matrix fully diagonal we need to eliminate nonzero entries on the row of position (t,jt) as well. This can be achieved by repeating the steps in Step II for columns instead of rows, and using multiplication on the right. In general this will result in the zero entries from the prior application of Step III becoming nonzero again. However, notice that the ideals generated by the elements at position (t,jt) form an ascending chain, because entries from a later step always divide entries from a previous step. Therefore, since R is a Noetherian ring (it is a PID), the ideals eventually become stationary and do not change. This means that at some stage after Step II has been applied, the entry at (t,jt) will divide all nonzero row or column entries before applying any more steps in Step II. Then we can eliminate entries in the row or column with nonzero entries while preserving the zeros in the already-zero row or column. At this point, only the block of A to the lower right of (t,jt) needs to be diagonalized, and conceptually the algorithm can be applied recursively, treating this block as a separate matrix. In other words, we can increment t by one and go back to Step I.

Smith normal form

12

Final step
Applying the steps described above to the remaining non-zero columns of the resulting matrix (if any), we get an -matrix with column indices where . The matrix entries are non-zero, and every other entry is zero. Now we can move the null columns of this matrix to the right, so that the nonzero entries are on positions . For short, set for the element at position . for which to , and then apply a row is a linear combination , and only: first add column The condition of divisibility of diagonal entries might not be satisfied. For any index one can repair this shortcoming by operations on rows and columns column to get an entry in column i without disturbing the entry equal to operation to make the entry at position of the original The value at position for

as in StepII; finally proceed as in

StepIII to make the matrix diagonal again. Since the new entry at position

, it is divisible by . does not change by the above operation (it is of the determinant of the upper

submatrix), whence that operation does diminish (by moving prime factors to the right) the value of

So after finitely many applications of this operation no further application is possible, which means that we have obtained as desired. Since all row and column manipulations involved in the process are invertible, this shows that there exist invertible and -matrices S, T so that the product S A T satisfies the definition of a Smith normal form. In particular, this shows that the Smith normal form exists, which was assumed without proof in the definition.

Applications
The Smith normal form is useful for computing the homology of a chain complex when the chain modules of the chain complex are finitely generated. For instance, in topology, it can be used to compute the homology of a simplicial complex or CW complex over the integers, because the boundary maps in such a complex are just integer matrices. It can also be used to prove the well known structure theorem for finitely generated modules over a principal ideal domain.

Example
As an example, we will find the Smith normal form of the following matrix over the integers.

The following matrices are the intermediate steps as the algorithm is applied to the above matrix.

Smith normal form So the Smith normal form is

13

and the elementary divisors are 2, 6 and 12.

Similarity
The Smith normal form can be used to determine whether or not matrices with entries over a common field are similar. Specifically two matrices A and B are similar if and only if the characteristic matrices and have the same Smith normal form. For example, with

A and B are similar because the Smith normal form of their characteristic matrices match, but are not similar to C because the Smith normal form of the characteristic matrices do not match.

References
Smith, Henry J. Stephen (1861). "On systems of linear indeterminate equations and congruences" [1]. Phil. Trans. R. Soc. Lond. 151 (1): 293326. doi:10.1098/rstl.1861.0016 [2]. Reprinted (pp. 367409 [3]) in The Collected Mathematical Papers of Henry John Stephen Smith, Vol. I [4], edited by J. W. L. Glaisher. Oxford: Clarendon Press (1894), xcv+603 pp. Smith normal form [5] at PlanetMath Example of Smith normal form [6] at PlanetMath K. R. Matthews, Smith normal form [7]. MP274: Linear Algebra, Lecture Notes, University of Queensland, 1991.

References
[1] [2] [3] [4] [5] [6] [7] http:/ / www. jstor. org/ stable/ 108738 http:/ / dx. doi. org/ 10. 1098%2Frstl. 1861. 0016 http:/ / archive. org/ stream/ collectedmathema01smituoft#page/ 366/ mode/ 2up http:/ / archive. org/ details/ collectedmathema01smituoft http:/ / planetmath. org/ encyclopedia/ GausssAlgorithmForPrincipalIdealDomains. html http:/ / planetmath. org/ encyclopedia/ ExampleOfSmithNormalForm. html http:/ / www. numbertheory. org/ courses/ MP274/ smith. pdf

Finitely-generated module

14

Finitely-generated module
In mathematics, a finitely generated module is a module that has a finite generating set. A finitely generated R-module also may be called a finite R-module or finite over R.[1] Related concepts include finitely cogenerated modules, finitely presented modules, finitely related modules and coherent modules all of which are defined below. Over a Noetherian ring the concepts of finitely generated, finitely related, finitely presented and coherent modules all coincide. A finitely generated module over a field is simply a finite-dimensional vector space, and a finitely generated module over the integers is simply a finitely generated abelian group.

Formal definition
The left R-module M is finitely generated if and only if there exist a1, a2, ..., an in M such that for all x in M, there exist r1, r2, ..., rn in R with x = r1a1 + r2a2 + ... + rnan. The set {a1, a2, ..., an} is referred to as a generating set for M in this case. In the case where the module M is a vector space over a field R, and the generating set is linearly independent, n is well-defined and is referred to as the dimension of M (well-defined means that any linearly independent generating set has n elements: this is the dimension theorem for vector spaces).

Examples
Let R be an integral domain with K its field of fractions. Then every R-submodule of K is a fractional ideal. If R is Noetherian, every fractional ideal arises in this way. Finitely generated modules over the ring of integers Z coincide with the finitely generated abelian groups. These are completely classified by the structure theorem, taking Z as the principal ideal domain. Finitely generated modules over division rings[citation needed] are precisely finite dimensional vector spaces.

Some facts
Every homomorphic image of a finitely generated module is finitely generated. In general, submodules of finitely generated modules need not be finitely generated. As an example, consider the ring R=Z[X1, X2, ...] of all polynomials in countably many variables. R itself is a finitely generated R-module (with {1} as generating set). Consider the submodule K consisting of all those polynomials with zero constant term. Since every polynomial contains only finitely many terms whose coefficients are non-zero, the R-module K is not finitely generated. In general, a module is said to be Noetherian if every submodule is finitely generated. A finitely generated module over a Noetherian ring is a Noetherian module (and indeed this property characterizes Noetherian rings): A module over a Noetherian ring is finitely generated if and only if it is a Noetherian module. This resembles, but is not exactly Hilbert's basis theorem, which states that the polynomial ring R[X] over a Noetherian ring R is Noetherian. Both facts imply that a finitely generated algebra over a Noetherian ring is again a Noetherian ring. More generally, an algebra (e.g., ring) that is a finitely-generated module is a finitely-generated algebra. Conversely, if a finitely generated algebra is integral (over the coefficient ring), then it is finitely generated module. (See integral element for more.) Let 0 M M M 0 be an exact sequence of modules. Then M is finitely generated if M, M are finitely generated. There are some partial converses to this. If M is finitely generated and M'' is finitely presented (which is stronger than finitely generated; see below), then M is finitely-generated. Also, M is Noetherian (resp. Artinian) if and only if M, M are Noetherian (resp. Artinian).

Finitely-generated module Let B be a ring and A its subring such that B is a faithfully flat right A-module. Then a left A-module F is finitely generated (resp. finitely presented) if and only if the B-module B A F is finitely generated (resp. finitely presented)[2].

15

Finitely generated modules over a commutative ring


For finitely generated modules over a commutative ring R, Nakayama's lemma is fundamental. Sometimes, the lemma allows one to prove finite dimensional vector spaces phenomena for finitely generated modules. For example, if f : M M is a surjective R-endomorphism of a finitely generated module M, then f is also injective, and hence is an automorphism of M.[3] This says simply that M is a Hopfian module. Similarly, an Artinian module M is coHopfian: any injective endomorphism f is also a surjective endomorphism[4]. Any R-module is an inductive limit of finitely generated R-submodules. This is useful for weakening an assumption to the finite case (e.g., the characterization of flatness with the Tor functor.) An example of a link between finite generation and integral elements can be found in commutative algebras. To say that a commutative algebra A is a finitely generated ring over R means that there exists a set of elements G = {x1, ..., xn} of A such that the smallest subring of A containing G and R is A itself. Because the ring product may be used to combine elements, more than just R combinations of elements of G are generated. For example, a polynomial ring R[x] is finitely generated by {1,x} as a ring, but not as a module. If A is a commutative algebra (with unity) over R, then the following two statements are equivalent[5]: A is a finitely generated R module. A is both a finitely generated ring over R and an integral extension of R.

Equivalent definitions and finitely cogenerated modules


The following conditions are equivalent to M being finitely generated (f.g.): For any family of submodules {Ni | i I} in M, if I. For any chain of submodules {Ni | i I} in M, if If is an epimorphism, then the restriction , then Ni = M for some i in I. is an epimorphism for some , then for some finite subset F of

finite subset F of I. From these conditions it is easy to see that being finitely generated is a property preserved by Morita equivalence. The conditions are also convenient to define a dual notion of a finitely cogenerated module M. The following conditions are equivalent to a module being finitely cogenerated (f.cog.): For any family of submodules {Ni | i I} in M, if of I. For any chain of submodules {Ni | i I} in M, if If is a monomorphism, then , then Ni = {0} for some i in I. is a monomorphism for some finite subset F , then for some finite subset F

of I. Both f.g. modules and f.cog. modules have interesting relationships to Noetherian and Artinian modules, and the Jacobson radical J(M) and socle soc(M) of a module. The following facts illustrate the duality between the two conditions. For a module M: M is Noetherian if and only if every submodule of N of M is f.g.

Finitely-generated module M is Artinian if and only if every quotient module M/N is f.cog. M is f.g. if and only if J(M) is a superfluous submodule of M, and M/J(M) is f.g. M is f.cog. if and only if soc(M) is an essential submodule of M, and soc(M) is f.g. If M is a semisimple module (such as soc(N) for any module N), it is f.g. if and only if f.cog. If M is f.g. and nonzero, then M has a maximal submodule and any quotient module M/N is f.g. If M is f.cog. and nonzero, then M has a minimal submodule, and any submodule N of M is f.cog. If N and M/N are f.g. then so is M. The same is true if "f.g." is replaced with "f.cog."

16

Finitely cogenerated modules must have finite uniform dimension. This is easily seen by applying the characterization using the finitely generated essential socle. Somewhat asymmetrically, finitely generated modules do not necessarily have finite uniform dimension. For example, an infinite direct product of nonzero rings is a finitely generated (cyclic!) module over itself, however it clearly contains an infinite direct sum of nonzero submodules. Finitely generated modules do not necessarily have finite co-uniform dimension either: any ring R with unity such that R/J(R) is not a semisimple ring is a counterexample.

Finitely presented, finitely related, and coherent modules


Another formulation is this: a finitely generated module M is one for which there is an epimorphism f : Rk M. Suppose now there is an epimorphism, : F M. for a module M and free module F. If the kernel of is finitely generated, then M is called a finitely related module. Since M is isomorphic to F/ker(), this basically expresses that M is obtained by taking a free module and introducing finitely many relations within F (the generators of ker()). If the kernel of is finitely generated and F has finite rank (i.e. F=Rk), then M is said to be a finitely presented module. Here, M is specified using finitely many generators (the images of the k generators of F=Rk) and finitely many relations (the generators of ker()). A coherent module M is a finitely generated module whose finitely generated submodules are finitely presented. Over any ring R, coherent modules are finitely presented, and finitely presented modules are both finitely generated and finitely related. For a Noetherian ring R, all four conditions are actually equivalent. Some crossover occurs for projective or flat modules. A finitely generated projective module is finitely presented, and a finitely related flat module is projective. It is true also that the following conditions are equivalent for a ring R: 1. R is a right coherent ring. 2. The module RR is a coherent module. 3. Every finitely presented right R module is coherent. Although coherence seems like a more cumbersome condition than finitely generated or finitely presented, it is nicer than them since the category of coherent modules is an abelian category, while, in general, neither finitely generated nor finitely presented modules form an abelian category.

Finitely-generated module

17

References
[1] [2] [3] [4] [5] For example, Matsumura uses this terminology. Bourbaki 1998, Ch 1, 3, no. 6, Proposition 11. Matsumura 1989, Theorem 2.4. Atiyah & Macdonald 1969, Exercise 6.1. Kaplansky 1970, p.11, Theorem 17.

Textbooks
Atiyah, M. F.; Macdonald, I. G. (1969), Introduction to commutative algebra, Addison-Wesley Publishing Co., Reading, Mass.-London-Don Mills, Ont., pp.ix+128, MR 0242802 (39 #4129) (http://www.ams.org/ mathscinet-getitem?mr=0242802+(39+#4129)) Bourbaki, Nicolas, Commutative algebra. Chapters 1--7. Translated from the French. Reprint of the 1989 English translation. Elements of Mathematics (Berlin). Springer-Verlag, Berlin, 1998. xxiv+625 pp. ISBN 3-540-64239-0 Kaplansky, Irving (1970), Commutative rings, Boston, Mass.: Allyn and Bacon Inc., pp.x+180, MR 0254021 (http://www.ams.org/mathscinet-getitem?mr=0254021) Lam, T. Y. (1999), Lectures on modules and rings, Graduate Texts in Mathematics No. 189, Springer-Verlag, ISBN978-0-387-98428-5 Lang, Serge (1997), Algebra (3rd ed.), Addison-Wesley, ISBN978-0-201-55540-0 Matsumura, Hideyuki (1989), Commutative ring theory, Cambridge Studies in Advanced Mathematics 8 (2 ed.), Cambridge: Cambridge University Press, pp.xiv+320, ISBN0-521-36764-6, MR 1011461 (90i:13001) (http:// www.ams.org/mathscinet-getitem?mr=1011461+(90i:13001)) Unknown parameter |note= ignored (help)

Free module
In mathematics, a free module is a free object in a category of modules. Given a set free module with basis . , a free module on is a Every vector space is free,[1] and the free vector space on a set is a special case of a free module on a set.

Definition
A free module is a module with a basis:[2] a linearly independent generating set. For an 1. 2. of ). If has invariant basis number, then by definition any two bases have the same cardinality. The cardinality of any , and is said to be free of rank n, or simply . are involved. (and therefore every) basis is called the rank of the free module free of finite rank if the cardinality is finite. Note that an immediate corollary of (2) is that the coefficients in (1) are unique for each The definition of an infinite free basis is similar, except that sum must still be finite, and thus for any particular In the case of an infinite basis, the rank of is the cardinality of only finitely many of the elements of . -module , the set ; for (where is the zero element of and distinct elements is the zero element of is a basis for if: is a finite sum of elements of multiplied

is a generating set for by coefficients in , then

; that is to say, every element of

is linearly independent, that is, if

will have infinitely many elements. However the

Free module

18

Construction
Given a set of Carrier: many) . , we define . Inverse: for , we define by , we define . A basis for is given by the set where by . Scalar multiplication: for by , we can construct a free contains the functions -module over . The module is simply the direct sum of copies , often denoted . We give a concrete realization of this direct sum, denoted by such that , as follows:

for cofinitely many (all but finitely

Addition: for two elements

(a variant of the Kronecker delta and a particular case of the indicator function, for the set Define the mapping vectors . by

). and the basis

. This mapping gives a bijection between

. We can thus identify these sets. Thus

may be considered as a linearly independent basis for

Universal property
The mapping such that defined above is universal in the following sense. If there is an arbitrary . -module and an arbitrary mapping , then there exists a unique module homomorphism

Generalisations
Many statements about free modules, which are wrong for general modules over rings, are still true for certain generalisations of free modules. Projective modules are direct summands of free modules, so one can choose an injection in a free module and use the basis of this one to prove something for the projective module. Even weaker generalisations are flat modules, which still have the property that tensoring with them preserves exact sequences, and torsion-free modules. If the ring has special properties, this hierarchy may collapse, i.e. for any perfect local Dedekind ring, every torsion-free module is flat, projective and free as well.

See local ring, perfect ring and Dedekind ring.

Free module

19

Notes
[1] Keown (1975), [2] Hazewinkel (1989),

References
Adamson, Iain T. (1972). Elementary Rings and Modules. University Mathematical Texts. Oliver and Boyd. pp.6566. ISBN0-05-002192-3. MR 0345993 (http://www.ams.org/mathscinet-getitem?mr=0345993). Keown, R. (1975). An Introduction to Group Representation Theory. Mathematics in science and engineering 116. Academic Press. ISBN978-0-12-404250-6. MR 0387387 (http://www.ams.org/ mathscinet-getitem?mr=0387387). Govorov, V. E. (2001), "Free module" (http://www.encyclopediaofmath.org/index.php?title=Free_module& oldid=13029), in Hazewinkel, Michiel, Encyclopedia of Mathematics, Springer, ISBN978-1-55608-010-4.

External links
This article incorporates material from free vector space over a set on PlanetMath, which is licensed under the Creative Commons Attribution/Share-Alike License.

Chinese remainder theorem


The Chinese Remainder Theorem is a result about congruences in number theory and its generalizations in abstract algebra. It was first published in the 3rd to 5th centuries by Chinese mathematician Sun Tzu. In its basic form, the Chinese remainder theorem will determine a number n that when divided by some given divisors leaves given remainders. For example, what is the lowest number n that when divided by 3 leaves a remainder of 2, when divided by 5 leaves a remainder of 3, and when divided by 7 leaves a remainder of 2? A common introductory example is a woman who tells a policeman that she lost her basket of eggs, and that if she makes three portions at a time out of it, she was left with 2, if she makes five portions at a time out of it, she was left with 3, and if she makes seven portions at a time out of it, she was left with 2. She then asks the policeman what is the minimum number of eggs she must have had. The answer to both problems is 23.

Theorem statement
The original form of the theorem, contained in the 5th-century book Sunzi's Mathematical Classic ( ) by the Chinese mathematician Sun Tzu and later generalized with a complete solution called Dayanshu ( ) in Qin Jiushao's 1247 Mathematical Treatise in Nine Sections ( , Shushu Jiuzhang), is a statement about simultaneous congruences. Suppose n1, n2, , nk are positive integers that are pairwise coprime. Then, for any given sequence of integers a1,a2, , ak, there exists an integer x solving the following system of simultaneous congruences.

Furthermore, all solutions x of this system are congruent modulo the product, N = n1n2nk. Hence for all , if and only if .

Chinese remainder theorem Sometimes, the simultaneous congruences can be solved even if the ni's are not pairwise coprime. A solution x exists if and only if:

20

All solutions x are then congruent modulo the least common multiple of the ni. Sun Tzu's work contains neither a proof nor a full algorithm. What amounts to an algorithm for solving this problem was described by Aryabhata (6th century; see Kak 1986). Special cases of the Chinese remainder theorem were also known to Brahmagupta (7th century), and appear in Fibonacci's Liber Abaci (1202). A modern restatement of the theorem in algebraic language is that for a positive integer with prime factorization we have the isomorphism between a ring and the direct product of its prime power parts:

Existence
Existence can be seen by an explicit construction of . We will use the notation inverse of coprime; the following construction explains why the coprimality condition is needed. to denote the multiplicative and are as calculated by the Extended Euclidean algorithm. It is defined exactly when

Case of two equations


Given the system (corresponding to )

Since

, we have from Bzout's identity

This is true because we agreed to use the inverses that came out of the Extended Euclidian algorithm; for any other inverses, it would not necessarily hold true, but only hold true . Multiplying both sides by , we get

If we take the congruence modulo

for the right-hand-side expression, it is readily seen that

But we know that

thus this suggests that the coefficient of the first term on the right-hand-side expression can be replaced by Similarly, we can show that the coefficient of the second term can be substituted by . We can now define the value

and it is seen to satisfy both congruences by reducing. For example

Chinese remainder theorem

21

General case
The same type of construction works in the general case of product of every modulus then define congruence equations. Let be the

and this is seen to satisfy the system of congruences by a similar calculation as before.

Finding the solution with basic algebra and modular arithmetic


For example, consider the problem of finding an integer x such that

A brute-force approach converts these congruences into sets and writes the elements out to the product of 345 = 60 (the solutions modulo 60 for each congruence): x {2, 5, 8, 11, 14, 17, 20, 23, 26, 29, 32, 35, 38, 41, 44, 47, 50, 53, 56, 59, } x {3, 7, 11, 15, 19, 23, 27, 31, 35, 39, 43, 47, 51, 55, 59, } x {1, 6, 11, 16, 21, 26, 31, 36, 41, 46, 51, 56, } To find an x that satisfies all three congruences, intersect the three sets to get: x {11, } Which can be expressed as

Another way to find a solution is with basic algebra, modular arithmetic, and stepwise substitution. We start by translating these congruences into equations for some t, s, and u: Equation 1: Equation 2: Equation 3: Start by substituting the x from equation 1 into congruence 2:

meaning that Plug t into equation 1:

for some integer s.

Plug this x into congruence 3:

Casting out fives, we get

meaning that

Chinese remainder theorem for some integer u. Finally,

22

So, we have solutions 11, 71, 131, 191, Notice that 60 = lcm(3,4,5). If the moduli are pairwise coprime (as they are in this example), the solutions will be congruent modulo their product.

A constructive algorithm to find the solution


The following algorithm only applies if the 's are pairwise coprime. (For simultaneous congruences when the moduli are not pairwise coprime, the method of successive substitution can often yield solutions.) Suppose, as above, that a solution is required for the system of congruences:

Again, to begin, the product For each ithe integers and such that and

is defined. Then a solution x can be found as follows. are coprime. Using the extended Euclidean algorithm we can find integers . Then, choosing the label , the above expression becomes:

Consider . The above equation guarantees that its remainder, when divided by , must be 1. On the other hand, since it is formed as , the presence of N guarantees a remainder of zero when divided by any when .

Because of this, and the multiplication rules allowed in congruences, one solution to the system of simultaneous congruences is:

For example, consider the problem of finding an integer x such that

Using the extended Euclidean algorithm, for x modulo 3 and 20 [45], we find (13) 3 + 2 20 = 1; i.e., e1 = 40. For x modulo 4 and 15 [35], we get (11) 4 + 3 15 = 1, i.e. e2 = 45. Finally, for x modulo 5 and 12 [34], we get 5 5 + (2) 12 = 1, i.e. e3 = 24. A solution x is therefore 2 40 + 3 45 + 1 (24) = 191. All other solutions are congruent to 191 modulo 60, [3 4 5 = 60], which means they are all congruent to 11 modulo 60. Note: There are multiple implementations of the extended Euclidean algorithm which will yield different sets of , , and . These sets however will produce the same solution; i.e., (20)2 + (15)3 + (24)1 = 109 = 11 modulo 60.

Chinese remainder theorem

23

Statement for principal ideal domains


For a principal ideal domain R the Chinese remainder theorem takes the following form: If u1, , uk are elements of R which are pairwise coprime, and u denotes the product u1uk, then the quotient ring R/uR and the product ring R/u1R R/ukR are isomorphic via the isomorphism such that

This map is well-defined and an isomorphism of rings; the inverse isomorphism can be constructed as follows. For each i, the elements ui and u/ui are coprime, and therefore there exist elements r and s in R with Set ei = s u/ui. Then the inverse of f is the map such that

This statement is a straightforward generalization of the above theorem about integer congruences: the ring Z of integers is a principal ideal domain, the surjectivity of the map f shows that every system of congruences of the form

can be solved for x, and the injectivity of the map f shows that all the solutions x are congruent modulo u.

Statement for general rings


The general form of the Chinese remainder theorem, which implies all the statements given above, can be formulated for commutative rings and ideals. If R is a commutative ring and I1, , Ik are ideals of R that are pairwise coprime (meaning that for all ), then the product I of these ideals is equal to their intersection, and the quotient ring R/I is isomorphic to the product ring R/I1 R/I2 R/Ik via the isomorphism such that

Here is a version of the theorem where R is not required to be commutative: Let R be any ring with 1 (not necessarily commutative) and canonical R-module homomorphism (as R-modules). is be pairwise coprime 2-sided ideals. Then the onto, with kernel . Hence,

Applications
In the RSA algorithm calculations are made modulo n, where n is a product of two large prime numbers p and q. 1,024-, 2,048- or 4,096-bit integers n are commonly used, making calculations in very time-consuming. By the Chinese remainder theorem, however, these calculations can be done in the isomorphic ring instead. Since p and q are normally of about the same size, that is about , calculations in the latter representation are much faster. Note that RSA algorithm implementations using this isomorphism are more susceptible to fault injection attacks. The Chinese remainder theorem may also be used to construct an elegant Gdel numbering for sequences, which is needed to prove Gdel's incompleteness theorems.

Chinese remainder theorem The following example shows a connection with the classic polynomial interpolation theory. Let r complex points ("interpolation nodes") be given, together with the complex data , for all and . The general Hermite interpolation problem asks for a polynomial each node : taking the prescribed derivatives in

24

Introducing the polynomials

the problem may be equivalently reformulated as a system of simultaneous congruences:

By the Chinese remainder theorem in the principal ideal domain with degree number case, can be performed as follows. Define the polynomials The partial fraction decomposition of gives r polynomials

, there is a unique such polynomial and such that .

. A direct construction, in analogy with the above proof for the integer

with degrees

so that

. Then a solution of the simultaneous congruence system is given by the polynomial

and the minimal degree solution is this one reduced modulo

, that is the unique with degree less than n.

The Chinese remainder theorem can also be used in secret sharing, which consists of distributing a set of shares among a group of people who, all together (but no one alone), can recover a certain secret from the given set of shares. Each of the shares is represented in a congruence, and the solution of the system of congruences using the Chinese remainder theorem is the secret to be recovered. Secret Sharing using the Chinese Remainder Theorem uses, along with the Chinese remainder theorem, special sequences of integers that guarantee the impossibility of recovering the secret from a set of shares with less than a certain cardinality. The Good-Thomas fast Fourier transform algorithm exploits a re-indexing of the data based on the Chinese remainder theorem. The Prime-factor FFT algorithm contains an implementation. Dedekind's theorem on the linear independence of characters states (in one of its most general forms) that if M is a monoid and k is an integral domain, then any finite family of distinct monoid homomorphisms (where the monoid structure on k is given by multiplication) is linearly independent; i.e., every family elements satisfying must be equal to the family . of

Proof using the Chinese Remainder Theorem: First, assume that k is a field (otherwise, replace the integral domain k by its quotient field, and nothing will change). We can linearly extend the monoid homomorphisms to k-algebra homomorphisms , where is the monoid ring of M over k. Then, the condition yields and and by linearity. Now, we notice that if and are two elements of the index set I, then the two k-linear maps (because if they were, then since every (since (since are not proportional to each other

would also be proportional to each other, and thus equal to each other are monoid homomorphisms), contradicting the assumption that they and are distinct. Now, is a field), and the ideals is a maximal ideal of and for are coprime whenever

be distinct). Hence, their kernels

(since they are distinct and maximal). The Chinese Remainder Theorem (for general rings) thus yields that the map

Chinese remainder theorem

25

given by for all is an isomorphism, where . Consequently, the map

given by for all is surjective. Under the isomorphisms , this map corresponds to the map

given by for every Now, QED. yields for every vector for every vector in the image of the map . Consequently, . Since is , surjective, this means that

Non-commutative case: a caveat


Sometimes in the commutative case, the conclusion of the Chinese Remainder Theorem is stated as . This version does not hold in the non-commutative case, since , as can be seen from the following example Consider the ring R of non-commutative real polynomials in x and y. Let I be the principal two-sided ideal generated by x and J the principal two-sided ideal generated by . Then but .

Proof
Observe that I is formed by all polynomials with an x in every term and that every polynomial in J vanishes under the substitution . Consider the polynomial . Clearly . Define a term in R as an element of the multiplicative monoid of R generated by x and y. Define the degree of a term as the usual degree of the term after the substitution . On the other hand, suppose . Observe that a term in q of maximum degree depends on y otherwise q under the substitution can not vanish. The same happens then for an element . Observe that the last y, from left to right, in a term of maximum degree in an element of is preceded by more than one x. (We are counting here all the preceding xs. E.g., in the last y is preceded by xs.) This proves that Hence . since that last y in a term of maximum degree ( ) is preceded by only one x.

On the other hand, it is true in general that implies . To see this, note that , while the opposite inclusion is obvious. Also, we have in general that, provided are pairwise coprime two-sided ideals in R, the natural map

is an isomorphism. Note that can be replaced by a sum over all orderings of (or just a sum over enough orderings, using inductively that for coprime ideals

of their product ).

Chinese remainder theorem

26

References
Donald Knuth. The Art of Computer Programming, Volume 2: Seminumerical Algorithms, Third Edition. Addison-Wesley, 1997. ISBN 0-201-89684-2. Section 4.3.2 (pp.286291), exercise 4.6.23 (page 456). Thomas H. Cormen, Charles E. Leiserson, Ronald L. Rivest, and Clifford Stein. Introduction to Algorithms, Second Edition. MIT Press and McGraw-Hill, 2001. ISBN 0-262-03293-7. Section 31.5: The Chinese remainder theorem, pp.873876. Laurence E. Sigler (trans.) (2002). Fibonacci's Liber Abaci. Springer-Verlag. pp.402403. ISBN0-387-95419-8. Kak, Subhash (1986), "Computational aspects of the Aryabhata algorithm" [1], Indian Journal of History of Science 21 (1): 6271. Thomas W. Hungerford (1974). Algebra. Springer-Verlag. pp.131132. ISBN0-387-90518-9. Cunsheng Ding, Dingyi Pei, and Arto Salomaa (1996). Chinese Remainder Theorem: Applications in Computing, Coding, Cryptography. World Scientific Publishing. pp.1213. ISBN981-02-2827-9.

External links
Hazewinkel, Michiel, ed. (2001), "Chinese remainder theorem" [2], Encyclopedia of Mathematics, Springer, ISBN978-1-55608-010-4 "Chinese Remainder Theorem" [3] by Ed Pegg, Jr., Wolfram Demonstrations Project, 2007. Weisstein, Eric W., "Chinese Remainder Theorem [4]", MathWorld. C# program and discussion [5] at codeproject University of Hawaii System [6] CRT by Lee Lady Full text of the Sunzi Suanjing [7] (Chinese) Chinese Text Project

References
[1] [2] [3] [4] [5] [6] [7] http:/ / www. ece. lsu. edu/ kak/ AryabhataAlgorithm. pdf http:/ / www. encyclopediaofmath. org/ index. php?title=p/ c022120 http:/ / demonstrations. wolfram. com/ ChineseRemainderTheorem/ http:/ / mathworld. wolfram. com/ ChineseRemainderTheorem. html http:/ / www. codeproject. com/ KB/ recipes/ CRP. aspx http:/ / www. math. hawaii. edu/ ~lee/ courses/ Chinese. pdf http:/ / ctext. org/ sunzi-suan-jing

Bzout's identity

27

Bzout's identity
Bzout's identity (also called Bezout's lemma) is a theorem in the elementary theory of numbers: let a and b be integers, not both zero, and let d be their greatest common divisor. Then there exist integers x and y such that

In addition, i) d is the smallest positive integer that can be written as ax + by, and ii) every integer of the form ax + by is a multiple of d. x and y are called Bzout coefficients for (a, b); they are not unique. A pair of Bzout coefficients (in fact the ones that are minimal in absolute value) can be computed by the extended Euclidean algorithm. Bzout's lemma is true in any principal ideal domain, but there are integral domains in which it is not true.

History
French mathematician tienne Bzout (17301783) proved this identity for polynomials.[1] However, this statement for integers can be found already in the work of another French mathematician, Claude Gaspard Bachet de Mziriac (15811638).[2][3][4]

Non-uniqueness of solutions
After one pair of Bzout coefficients (x, y) has been computed (using extended Euclid or some other algorithm), all pairs may be found using the formula

Example
Let a = 12 and b = 42, gcd(12, 42) = 6. Then

Generalizations
Bzout's identity can be extended to more than two integers: if

then there are integers such that

and 1) d is smallest positive integer of this form, and 2) every number of this form is a multiple of d. As noted in the introduction, Bzout's identity works not only in the ring of integers, but also in any other principal ideal domain (PID). That is, if R is a PID, and a and b are elements of R, and d is a greatest common divisor of a and b, then there are elements x and y in R such that ax + by = d. The reason: the ideal Ra+Rb is principal and indeed is

Bzout's identity equal to Rd. An integral domain in which Bzout's identity holds is called a Bzout domain.

28

Proof
Bzout's lemma is a consequence of the Euclidean division defining property, namely that the division by a nonzero integer b has a remainder strictly less than |b|. The proof that follows may be adapted for any Euclidean domain. For given nonzero integers a and b there is a nonzero integer d = as + bt of minimal absolute value among all those of the form ax + by with x and y integers; one can assume d > 0 by changing the signs of both s and t if necessary. Now the remainder of dividing either a or b by d is also of the form ax + by since it is obtained by subtracting a multiple of d = as + bt from a or b, and on the other hand it has to be strictly smaller in absolute value than d. This leaves 0 as only possibility for such a remainder, so d divides a and b exactly. If c is another common divisor of a and b, then c also divides as + bt = d. Since c divides d but is not equal to it, it must be less than d. This means that d is the greatest common divisor of a and b; this completes the proof.

Notes
[1] Bzout, Thorie gnrale des quations algbriques (http:/ / books. google. fr/ books?id=FoxbAAAAQAAJ& hl=en& pg=PP5#v=onepage& q& f=false) (Paris, France: Ph.-D. Pierres, 1779). [3] On these pages, Bachet proves (without equations) Proposition XVIII. Deux nombres premiers entre eux estant donnez, treuver le moindre multiple de chascun diceux, surpassant de lunit un multiple de lautre. (Given two numbers [which are] relatively prime, find the lowest multiple of each of them [such that] one multiple exceeds the other by unity (1).) This problem (namely, ax - by = 1) is a special case of Bzouts equation and was used by Bachet to solve the problems appearing on pages 199 ff. [4] See also:

External links
Online calculator (http://wims.unice.fr/wims/wims.cgi?module=tool/arithmetic/bezout.en) of Bzout's identity. Weisstein, Eric W., " Bezouts Identity (http://mathworld.wolfram.com/BezoutsIdentity.html)", MathWorld.

Article Sources and Contributors

29

Article Sources and Contributors


Structure theorem for finitely generated modules over a principal ideal domain Source: https://en.wikipedia.org/w/index.php?oldid=549870933 Contributors: Alecobbe, Algebraist, Altrevolte, Anne Bauval, ArnoldReinhold, Arthur Rubin, Auntof6, Bomazi, Brad7777, David Eppstein, EmilJ, Expz, Foobarnix, Foxjwill, Frap, Geometry guy, Giftlite, Grafen, Henning Makholm, JackSchmidt, John a s, Kastberg, Kundor, Marc van Leeuwen, Mct mht, Michael Hardy, Mulanhua, Nbarth, Oleg Alexandrov, Ozob, ReyBrujo, Rich Farmbrough, Roentgenium111, Rschwieb, Silly rabbit, Tobias Bergemann, Vincent Semeria, Zelmerszoetrop, Zundark, 19 anonymous edits Torsion (algebra) Source: https://en.wikipedia.org/w/index.php?oldid=567314197 Contributors: Anita5192, Arcfrk, Bhny, Bia87321, Charles Matthews, Chemturion, DavidCBryant, Delaszk, Deltahedron, Giftlite, JackSchmidt, Jitse Niesen, MathMartin, Michael Hardy, Number 0, PatrickR2, Quondum, R.e.b., Rjwilmsi, Rschwieb, Sam Derbyshire, Sct72, Silly rabbit, TakuyaMurata, Tesseran, Tobias Bergemann, Zundark, 21 anonymous edits Zero divisor Source: https://en.wikipedia.org/w/index.php?oldid=566860508 Contributors: 62.100.0.xxx, Andres, Astrale, Atlantia, AxelBoldt, Bryan Derksen, CRGreathouse, Carvasf, Chricho, CiaPan, Conversion script, Dan Gardner, Eleuther, Eric119, Ersik, Etale, Feinoha, GaborLajos, Giftlite, Gleuschk, Graham87, Helder.wiki, Incnis Mrsi, Jarble, Jeepday, JoergenB, Joseph Dwayne, Jshadias, Karl-Henner, Lambiam, Little green rosetta, Magioladitis, MathMartin, MathyGuy23, Mav, Melchoir, Michael Hardy, Michael Slone, OneWeirdDude, Policron, Quondum, RDBury, Rschwieb, Sabbut, Sabin4232, Savantas83, Sheogorath, SirJective, TakuyaMurata, Tobias Bergemann, Trumpet marietta 45750, Vivacissamamente, Xantharius, Zhongmou Zhang, 39 anonymous edits Smith normal form Source: https://en.wikipedia.org/w/index.php?oldid=570674310 Contributors: Andreas Carter, Aquae, Bekant, Benzi455, Charles Matthews, David Eppstein, Dysprosia, FlyHigh, Gauge, Giftlite, GromXXVII, InverseHypercube, Keith Lehwald, Marc van Leeuwen, Maxal, Michael Hardy, Nbarth, PappyK, Radagast83, Rgdboer, Schickling, Seanwal111111, Szquirrel, TakuyaMurata, The Anome, The Rambling Man, Tomonacci, Zundark, 26 anonymous edits Finitely-generated module Source: https://en.wikipedia.org/w/index.php?oldid=562931402 Contributors: AHusain314, Agraboso, AxelBoldt, B. Wolterding, Barticus88, Charles Matthews, Commentor, Danneks, Dysprosia, ElNuevoEinstein, Frap, HarDan, Helohe, JoergenB, JumpDiscont, Michael Hardy, Number 0, Onebyone, Ozob, Pfortuny, Quondum, RobHar, Rschwieb, Stickee, Sawomir Biay, TakuyaMurata, Vincent Semeria, Waltpohl, Wavelength, 8 anonymous edits Free module Source: https://en.wikipedia.org/w/index.php?oldid=561721504 Contributors: 16@r, 777sms, Algebraist, Beroal, Certes, Charles Matthews, ComplexZeta, David Eppstein, DefLog, Dmharvey, Esqg, Fropuff, Giftlite, Helder.wiki, Hillman, JackSchmidt, Jitse Niesen, JoergenB, KonradVoelkel, Lethe, Linas, Maksim-e, Martha Van Della, MathMartin, Nbarth, Oleg Alexandrov, Onebyone, Rodhullandemu, Salix alba, TakuyaMurata, Trovatore, Vanish2, Waltpohl, 20 anonymous edits Chinese remainder theorem Source: https://en.wikipedia.org/w/index.php?oldid=569877752 Contributors: 2001:A60:F0E5:0:5CC9:4E37:BFF2:158A, 62.100.20.xxx, Alcazar84, Algebraist, Andrei Stroe, Arcfrk, Arthur Rubin, AxelBoldt, Bazonka, BeteNoir, Black Carrot, Bluefoxicy, Bryan Derksen, CRGreathouse, CYD, CaoWiki, Charles Matthews, Charleswallingford, Cikicdragan, Connor Behan, Conversion script, CryptoBm, DMacks, DVdm, Da Joe, Darij, David Battle, David Eppstein, Dcoetzee, DemonThing, Die Mensch-Maschine, Dionyziz, Dmharvey, DonDiego, Doomhydra1098, DrTJJ, Dratman, Drostie, Druckles, Dysprosia, El C, Error792, Fenris.kcf, Fly by Night, Franklin.vp, Fuzheado, Garald, Georg Muntingh, Giftlite, Gisling, Glrx, Goochelaar, Guardian of the Rings, Heilmoli, Jannaston, Jebba, Jj137, Jmcc150, JoergenB, KSmrq, Kallikanzarid, Kesla, Kompik, Kurykh, LlywelynII, Lotje, Lowellian, Lupin, MC-CPO, MSGJ, Manoguru, Matthew Woodcraft, Meekohi, Mendevel, Mentifisto, Michael Hardy, Michael Slone, Mindmatrix, Mjaiclin, Mpatel, Nbarth, Newone, Nickj, Nilram12, Ojigiri, Oleg Alexandrov, PMajer, PericlesofAthens, PierreAbbat, Plclark, Pleasantville, Policron, Preetum, Protonk, Python eggs, Quandle, Quanticle, Qwertyca, RDBury, Ranjith1610, Rbb l181, Reedy, Salgueiro, Sapphic, Scott Ritchie, Shahab, Stevertigo, Stuart Morrow, Taral, Taw, Thomasyen, Trusilver, Typometer, Vincent Semeria, WLior, Wykypydya, X127, Your mind17, Zeno Gantner, Zundark, 175 anonymous edits Bzout's identity Source: https://en.wikipedia.org/w/index.php?oldid=569275921 Contributors: Anghammarad, AxelBoldt, Azuredu, Bender235, Blokhead, Bryan Derksen, Burn, Caarecengi, Calle, CarlFriedrich, Charles Matthews, Conversion script, Cwkmail, D.Lazard, David Eppstein, David Haslam, Dcirovic, Dupuju, Gandalf61, Garfieldnate, Gene Nygaard, Giftlite, Grover cleveland, Hanche, Heron, Icarot, Ilya Voyager, JadeNB, JanKG, Jitse Niesen, Justin W Smith, Keenan Pepper, Kier07, Kompik, Konradek, LGB, Marc van Leeuwen, Materialscientist, MathMartin, Maxal, Merovingian, Michael Hardy, Michael Ross, Nd12345, Ntsimp, Ohanian, Oleg Alexandrov, Oliver Pereira, Peterepeat11, Ramanujan srinivasa, RedAcer, Rich Farmbrough, Rick Ballan, Sadeq, Sanderling, Shreevatsa, Spoon!, Stdazi, TheKing44, Timwi, Tuntable, VKokielov, Virginia-American, Wmahan, Wtt, Xmlizer, Zundark, Zvika, 37 anonymous edits

Image Sources, Licenses and Contributors

30

Image Sources, Licenses and Contributors


Image:Lattice torsion points.svg Source: https://en.wikipedia.org/w/index.php?title=File:Lattice_torsion_points.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors: User:Sam Derbyshire File:Module properties in commutative algebra.svg Source: https://en.wikipedia.org/w/index.php?title=File:Module_properties_in_commutative_algebra.svg License: Creative Commons Zero Contributors: User:KonradVoelkel

License

31

License
Creative Commons Attribution-Share Alike 3.0 Unported //creativecommons.org/licenses/by-sa/3.0/

You might also like