Yan Sheng's site A math blog

Commutative Algebra (III): Finite Modules

Let AA be a ring. An AA-module MM is an abelian group equipped with a scalar multiplication A×MMA\times M\to M, sending (a,m)(a,m) to ama\cdot m (usually written amam), satisfying associativity:

(aa)m=a(am),(aa')m=a(a'm),

distributivity:

a(m+m)=am+am,(a+a)m=am+am,a(m+m')=am+am',\quad(a+a')m=am+a'm,

and 1m=m1m=m.

If we fix aAa\in A, then scalar multiplication by aa gives a map mamm\mapsto am which is a (group) endomorphism of MM. Moreover, if we consider the (non-commutative) endomorphism ring End(M)\End(M), with addition defined pointwise and multiplication given by composition, then an AA-module structure on MM is equivalent to a ring homomorphism AEnd(M)A\to\End(M). This is sometimes called the AA-module structure homomorphism.

This alternative viewpoint is sometimes more convenient. For instance, if MM is a BB-module and f:ABf:A\to B is a ring homomorphism, then MM gets an AA-module structure if we define am=f(a)ma\cdot m=f(a)m. We can verify this fact by explicitly checking all module axioms, or we could simply note that the composition of ff with the BB-module structure homomorphism BEnd(M)B\to\End(M) gives a ring homomorphism AEnd(M)A\to\End(M), which must then define an AA-module structure.

The determinant trick

We say that an AA-module MM is finite (over AA) if MM is finitely generated as an AA-module. Finite AA-modules are amenable to the following important “determinant trick,” which yields a generalisation of the Cayley–Hamilton theorem in linear algebra:

Proposition: Let MM be a finite AA-module generated by nn elements, and let φHomA(M,M)\varphi\in\Hom_A(M,M). If IAI\subseteq A is an ideal such that φ(M)IM\varphi(M)\subseteq IM, then there exists aiIia_i\in I^i such that

φn+a1φn1++an1φ+an=0,\varphi^n+a_1\varphi^{n-1}+\cdots+a_{n-1}\varphi+a_n=0,

as endomorphisms of MM.

Proof: Let M=Aω1++AωnM=A\omega_1+\cdots+A\omega_n. We will first write the condition φ(M)IM\varphi(M)\subseteq IM as a “matrix equation”: for each ii, we have

φ(ωi)=j=1naijωj\varphi(\omega_i)=\sum_{j=1}^na_{ij}\omega_j

for some aijIa_{ij}\in I. Moving terms to one side, we get the following for all ii:

j=1n(δijφaijIdM)ωj=0,\sum_{j=1}^n(\delta_{ij}\varphi-a_{ij}\Id_M)\omega_j=0,

where δij\delta_{ij} is the Kronecker delta, and δijφaijIdM\delta_{ij}\varphi-a_{ij}\Id_M is interpreted as an endomorphism of MM.

The next step is to consider the determinant of the matrix (δijφaijIdM)i,j(\delta_{ij}\varphi-a_{ij}\Id_M)_ {i,j}. As it stands, this is not a well-defined concept, since End(M)\End(M) is not necessarily commutative.

However, let us consider the subring A[φ]End(M)A[\varphi]\subseteq\End(M) generated by AIdMA\Id_M and φ\varphi; this is the collection of AA-linear combinations of powers of φ\varphi. Since AA-multiplication commutes with all endomorphisms of MM, and powers of φ\varphi commute, the ring A[φ]A[\varphi] is in fact commutative, and we can proceed.

The matrix (δijφaijIdM)i,j(\delta_{ij}\varphi-a_{ij}\Id_M)_ {i,j} is a matrix with coefficients in A[φ]A[\varphi], so it has well-defined determinant dd and cofactors bijb_{ij}. Multiplying both sides of the matrix equation by the cofactor matrix, we get

dωk=0for all k.d\omega_k=0\quad\text{for all }k.

Hence dM=0dM=0 implies d=0d=0 as an element of End(M)\End(M). Now expanding the determinant dd gives us a relation of the desired form.  \qed

Note that from the appearance of the characteristic polynomial in the proof, we can recover the usual Cayley–Hamilton theorem when MM is a free AA-module and I=AI=A.

Nakayama’s lemma

The next result is also commonly attributed to Krull–Azumaya, and is an important result throughout commutative algebra.

Nakayama’s Lemma: Let MM be a finite AA-module, and let IAI\subseteq A be an ideal such that M=IMM=IM. Then there exists aAa\in A such that aM=0aM=0 and a1(modI)a\equiv1\pmod I.

In particular, if Irad(A)I\subseteq\rad(A) then M=0M=0.

Proof: Set φ=IdM\varphi=\Id_M in the previous proposition, to get

a=1+a1++an=0a=1+a_1+\cdots+a_n=0

as endomorphisms of MM, so aM=0aM=0. Furthermore, aiIa_i\in I for all ii implies a1(modI)a\equiv1\pmod I.

If Irad(A)I\subseteq\rad(A), then a1(modrad(A))a\equiv1\pmod{\rad(A)} implies that aa is a unit of AA, so M=a1(aM)=0M=a^{-1}(aM)=0.  \qed

Proving that a module is equal to zero might not seem impressive at first sight, but many important properties in module theory can be rephrased in these terms. For instance, a submodule NMN\subseteq M is equal to the whole module if and only if M/N=0M/N=0; a module homomorphism is injective if and only if its kernel is zero, and surjective if and only if its cokernel is zero; and so on. We will see some examples of this in the next section.

We can restate Nakayama’s lemma as follows.

Corollary: Let MM be a finite AA-module, and let IAI\subseteq A be an ideal such that M=IMM=IM. Then there exists bIb\in I such that bm=mbm=m for all mMm\in M.

Proof: Let aa be as in Nakayama’s lemma, and take b=1ab=1-a.  \qed

In other words, if M=IMM=IM then some element of II is an obvious witness, by acting as identity on MM.

The most important applications of Nakayama’s lemma are to local rings:

Corollary: Let (A,m,k)(A,\mf m,k) be a local ring, and MM be a finite AA-module. If M/mM=0M/\mf mM=0, then M=0M=0.  \qed

Hence the problem of showing that the AA-module MM is zero is reduced to showing that the kk-vector space M/mMM/\mf mM is zero, for which we have the well-developed tools of linear algebra at our disposal.

Applications of Nakayama’s lemma

Proposition: Let MM be an AA-module, and NMN\subseteq M an AA-submodule such that M/NM/N is finite over AA. If Irad(A)I\subseteq\rad(A) is an ideal of AA such that M=N+IMM=N+IM, then M=NM=N.

Proof: Let M=M/N\overline M=M/N. Then the given condition is equivalent to M=IM\overline M=I\overline M, so Nakayama’s lemma gives M=0\overline M=0, ie. M=NM=N.  \qed

A minimal basis of an AA-module MM is a set WMW\subseteq M which generates MM, such that no proper subset of WW generates MM; in other words, a minimal generating set.

For vector spaces, the notions of minimal generating sets and maximal linearly independent sets are equivalent, but this is not true for arbitrary modules. For instance, viewing Q\bb Q as a Z\bb Z-module, no finite subset is generating, so any minimal basis is infinite; but no two elements are Z\bb Z-linearly independent, so maximal linearly independent sets have only one element!

Also, note that two minimal bases need not have the same number of elements; for instance, if M=AM=A, and (x),(y)(x),(y) are two coprime ideals, then {1}\{1\} and {x,y}\{x,y\} are both minimal bases of AA.

However, for local rings we have the following result:

Proposition: Let (A,m,k)(A,\mf m,k) be a local ring, and MM be a finite AA-module. Let M=M/mM\overline M=M/\mf mM, which is a kk-vector space, and let n=dimkMn=\dim_k\overline M.

  1. For any basis {u1,,un}\{\overline u_1,\ldots,\overline u_n\} of M\overline M over kk, choose inverse images uiMu_i\in M for each ui\overline u_i. Then {u1,,un}\{u_1,\ldots,u_n\} is a minimal basis of MM.
  2. Conversely, for every minimal basis {u1,,un}\{u_1,\ldots,u_{n’}\}, the images {u1,,un}\{\overline u_1,\ldots,\overline u_{n’}\} form a basis of M\overline M. Hence n=nn’=n, ie. every minimal basis has nn elements.
  3. If {u1,,un}\{u_1,\ldots,u_n\} and {v1,,vn}\{v_1,\ldots,v_n\} are two minimal bases of MM, and vi=jaijujv_i=\sum_ja_{ij}u_j with aijAa_{ij}\in A, then det(aij)\det(a_{ij}) is a unit of AA, so (aij)(a_{ij}) is an invertible matrix.

Proof: (1) Note that M=Aui+mMM=\sum Au_i+\mf mM, and MM is finite implies M/AuiM/\sum Au_i is finite. Hence by the previous proposition, M=AuiM=\sum Au_i.

Now if {u1,,un}\{u_1,\ldots,u_n\} is not minimal, say {u1,,un1}\{u_1,\ldots,u_{n-1}\} generates MM, then {u1,,un1}\{\overline u_1,\ldots,\overline u_{n-1}\} generates M\overline M, contradiction. Hence {u1,,un}\{u_1,\ldots,u_n\} is a minimal basis of MM.

(2) Since {u1,,un}\{u_1,\ldots,u_{n’}\} generates MM, we have {u1,,un}\{\overline u_1,\ldots,\overline u_{n’}\} generates M\overline M.

Now if there is a linear relation among {u1,,un}\{\overline u_1,\ldots,\overline u_{n’}\}, then some proper subset generates M\overline M; by (1), this corresponds to a proper subset of {u1,,un}\{u_1,\ldots,u_{n’}\} which generates MM, contradiction. Hence {u1,,un}\{\overline u_1,\ldots,\overline u_{n’}\} is linearly independent, so it is a basis of M\overline M.

(3) Let aij\overline a_{ij} be the image of aija_{ij} in k=A/mk=A/\mf m, so vi=aijuj\overline v_i=\sum\overline a_{ij}\overline u_j. Now (aij)(\overline a_{ij}) represents a change of basis over M\overline M, so

0det(aij)=det(aij)(modm),0\neq\det(\overline a_{ij})=\det(a_{ij})\pmod{\mf m},

so det(aij)\det(a_{ij}) is a unit of AA. Then the inverse of (aij)(a_{ij}) is given by Cramer’s formula.  \qed

Proposition: Let MM be a finite AA-module. If f:MMf:M\to M is a surjective AA-linear map, then ff is also injective (so it is an automorphism of MM).

Proof: We will need the universal property of A[X]A[X] over noncommutative rings: if RR is a (not necessarily commutative) ring, φ:AR\varphi:A\to R is a ring homomorphism, and rRr\in R is an element commuting with every φ(a)\varphi(a), then there exists a ring homomorphism φ~:A[X]R\tilde\varphi:A[X]\to R extending φ\varphi and sending XX to rr.

Now take R=End(M)R=\End(M), φ:AEnd(M)\varphi:A\to\End(M) the structure homomorphism defining the AA-module structure on MM, and r=fr=f. Then the lift φ~:A[X]End(M)\tilde\varphi:A[X]\to\End(M) defines an A[X]A[X]-module structure on MM, where Xm=f(m)X\cdot m=f(m).

By assumption, we have XM=MXM=M, ie. (X)M=M(X)M=M, so by Nakayama’s lemma there exists XP(X)(X)XP(X)\in(X) acting as identity on MM. Hence φ~(P(X))End(M)\tilde\varphi(P(X))\in\End(M) is an inverse of ff, so ff is injective.  \qed

Comments (0)

Be the first to comment!

Write a comment…