Section 16.1 Depth and Grade
Definition 16.1.
Let \(R\) be a ring, \(M\) an \(R\)-module, and \({\bf x}=x_1,\dots,x_n\in R\text{.}\) We say \({\bf x}\) is a regular sequence on \(M\), or simply an \(M\)-sequence, if \((x_1,\dots, x_n)M\neq M\) and \(x_i\) is regular (i.e., a non-zero-divisor) on \(M/(x_1,\dots,x_{i-1})M\) for each \(i=1,\dots,n\text{.}\) An \(M\)-sequence \({\bf x}\) is called maximal if \({\bf x}\) cannot be extended to a longer \(M\)-sequence.
Example 16.2.
Example 49. Let \(R=k[x_1,\dots,x_n]\) be a polynomial ring over the field \(k\text{.}\) Then \(x_1,\dots,x_n\) is a maximal \(R\)-sequence.
Lemma 16.3.
Let \(R\) be Noetherian and \(M\) a finitely generated \(R\)-module. Then
The length of every \(M\)-sequence is finite.
Every \(M\)-sequence can be extended to a maximal \(M\)-sequence.
Proof.
For part (a), note that if \(x_1,\dots,x_n\) is an \(M\)-sequence then
\begin{equation*}
0\subset (x_1)M\subset (x_1,x_2)M\subset \cdots \subset (x_1,\dots,x_n)M
\end{equation*}
is a strictly ascending chain of submodule of \(M\text{.}\) For if \((x_1,\dots,x_i)M=(x_1,\dots,x_{i+1})M\) for some \(i\text{,}\) then \(x_{i+1}M\subseteq (x_1,\dots,x_i)M\text{.}\) As \(M\neq (x_1,\dots,x_i)M\text{,}\) then \(x_{i+1}\) is a zero-divisor on \(M/(x_1,\dots,x_i)M\text{,}\) a contradiction. Consequently, as \(M\) is Noetherian, this chain must terminate. Hence, there are no \(M\)-sequences of infinite length. The same argument (using ACC) proves (b) as well.
Example 16.4.
Let \(R=k[x,y,x]\text{.}\) Then \(x,y,z\) is a maximal \(R\)-sequence.
Example 16.5.
Let \(R=k[x,y,z]_{(x,y,z)}\) and \(M=R/(xy,yz)\text{.}\) Then \(y-z\) is a maximal \(M\) -sequence. To see this, note that \((xy,xz)=(y)\cap (x,z)\) is a irredundant primary decomposition of \(I\text{.}\) Thus, \((y)\) and \((x,z)\) are the associated primes of \(M\text{.}\) Since \(y-z\) is not in either associated prime, it is a non-zero-divisor on \(M\text{.}\) Also, \(M\neq (y-z)M\text{,}\) so \(y-z\) is \(M\)-regular. Note that \(M/(y-z)M\cong k[x,y,z]_{(x,y,z)}/(xy, xz, y-z)\cong k[x,y]_{(x,y)}/(xy, y^2)\text{.}\) Since \((x,y)x=0\) in \(k[x,y]_{(x,y)}/(xy, y^2)\text{,}\) \((x,y)\) consists of zero-divisors on \(M/(y-z)M\text{.}\) Thus, \(y-z\) is a maximal \(M\)-sequence.
Here we summarize some essential facts about primary decompositions for modules, and then use them to prove Krull’s Intersection Theorem.
Definition 16.6.
Let \(M\) be an \(R\)-module. A submodule \(Q\) of \(M\) is called primary if \(Q\subsetneq M\) and for every \(r\in R\) multiplication by \(r\) on \(M/Q\) is either injective or nilpotent; i.e., \(r\) is either a non-zero-divisor on \(M/Q\) or \(r^nM\subseteq Q\) for some \(n\text{.}\)
Theorem 16.8.
Let \(R\) be a Noetherian ring, \(M\) a finitely generated \(R\)-module and \(N\subsetneq M\) a submodule. Then there exists primary submodules \(Q_1,\dots,Q_n\) such that
\(N=Q_1\cap \cdots \cap Q_n\text{;}\)
\(N\subsetneq Q_1\cap \cdots \cap \hat{Q_i}\cap \cdots Q_n\) for \(i=1,\dots,n\text{;}\)
\(p_1,\dots,p_n\) are distinct prime ideals, where \(p_i=\sqrt{\operatorname{Ann}_R M/Q_i}\text{.}\)
The decomposition \(N=Q_1\,\cap\cdots\cap\, Q_n\) is called an irredundant primary decomposition for \(N\subset M\). The prime ideals \(p_1,\dots,p_n\) are uniquely determined by \(N\subset M\) and are called the associated primes of \(N\subset M\) (or more commonly, of \(M/N\)). We denote the set of associated primes of \(M/N\) by \(\operatorname{Ass}_R M/N\text{.}\) Moreover, a prime ideal \(p\in \operatorname{Ass}_R M/N\) if and only if \(p=(N:_R x)\) for some \(x\in M\text{.}\) Additionally, if \(p_i\) is a minimal associated prime of \(M/N\text{,}\) the \(Q_i=\phi^{-1}(N_{p_i})\) where \(\phi:M\to M_{p_i}\) is the natural localization map.
Proof.
See Atiyah-Macdonald. ◻
Theorem 16.9. Krull’s Intersection Theorem.
Let \(R\) be a Noetherian ring, \(I\) an ideal, and \(M\) a finitely generated \(R\)-module. Then there exist an \(s\in I\) such that
\begin{equation*}
(1-s)\bigcap_{n=1}^{\infty} I^nM=0.
\end{equation*}
Proof.
Proof. Let \(\displaystyle{N=\bigcap_{n=1}^{\infty} I^nM}\text{.}\) We claim that \(IN=N\text{.}\) If \(IN=M\text{,}\) there is nothing to prove. Suppose \(IN\subsetneq M\) and let \(IN=Q_1\cap Q_2\cap \cdots \cap Q_r\) be a primary decomposition of \(IN\subset M\text{.}\) Then for each \(i\text{,}\) \(IN\subseteq Q_i\text{.}\) If \(N\not\subset Q_i\) then \(I\) consists of zero-divisors on \(M/Q_i\text{.}\) As \(Q_i\) is a primary submodule of \(M\text{,}\) we must have \(I^nM\subseteq Q_i\) for some \(n\text{.}\) But \(N\subseteq I^nM\text{,}\) so \(N\subseteq Q_i\text{,}\) a contradiction. Thus, \(N\subseteq Q_1\cap\cdots \cap Q_r=IN\text{.}\) Consequently, \(IN=N\text{.}\) By a homework exercise, this implies there exists \(s\in I\) such that \((1-s)N=0\text{.}\) ◻
Definition 16.10.
Let \(R\) be a ring. The Jacobson radical of \(R\), denoted \(J(R)\text{,}\) is the intersection of all maximal ideals of \(R\text{.}\) It is easily seen that if \(r\in J(R)\) then \(1-r\) is a unit.
Corollary 16.11.
Let \(R\) be a Noetherian ring, \(I\subseteq J(R)\) an ideal, and \(M\) a finitely generated \(R\)-module. Then
\begin{equation*}
\bigcap_{n=1}^{\infty} I^nM=0.
\end{equation*}
Proof.
Apply Krull’s Intersection Theorem and use that \(1-s\) is a unit for every \(s\in I\text{.}\)
Proposition 16.12.
Let \(R\) be Noetherian, \(M\) a finitely generated \(R\)-module, and \({\bf x}=x_1,\dots,x_n\in J(R)\) an \(M\)-sequence. Then any permutation of \({\bf x}\) is an \(M\)-sequence.
Proof.
It suffices to show that if \(x,y\in J(R)\) is an \(M\)-sequence then so is \(y,x\text{.}\) First note that \((y,x)M=(x,y)M\neq M\text{.}\) We next show that \(y\) is regular on \(M\text{:}\) suppose \(yu=0\) for some \(u\in M\text{.}\) Then \(yu\in xM\) so \(u\in xM\text{.}\) Write \(u=xu_1\) where \(u_1\in M\text{.}\) Then \(0=yu=xyu_1\text{,}\) so \(yu_1=0\text{.}\) Repeating the argument above, we get \(u_1\in xM\) and hence \(u\in (x)^2M\text{.}\) Continuing in this way, we obtain that \(u\in \bigcap_{n\geqslant 1} (x)^nM\text{.}\) As \((x)\subseteq J(R)\) we have \(\bigcap_{n\geqslant 1} (x)^nM=0\) and thus \(u=0\text{.}\) This shows \(y\) is regular on \(M\text{.}\) Now assume \(xv\in yM\) for some \(v\in M\text{.}\) Then \(xv=yw\) for some \(w\in M\text{.}\) Since \(y\) is regular on \(M/xM\text{,}\) we obtain that \(w=xz\) for some \(z\in M\text{.}\) Consequently, \(xv=xyz\text{.}\) As \(x\) is regular on \(M\text{,}\) we then have \(v=yz\in yM\text{.}\) Hence, \(x\) is regular on \(M/yM\text{.}\)
Definition 16.13.
Let \(R\) be a Noetherian ring, \(I\) an ideal of \(R\text{,}\) and \(M\) a finitely generated \(R\)-module such that \(IM\neq M\text{.}\) Then the grade of \(I\) on \(M\), denote \(\operatorname{grade}_I M\) or \(\operatorname{grade}(I,M)\text{,}\) is defined to be the supremum of the lengths of all \(M\)-sequences contained in \(I\text{.}\)
Proposition 16.16.
Let \(R\) be a ring, \(L\) and \(M\) \(R\)-modules, and \(x\in R\) such that \(xL=0\) and \(x\) is regular on \(M\text{.}\) Then \(g(L,M/xM)=g(L,M)-1\text{.}\) Furthermore, if \(g:=g(L,M)<\infty\) then \(\operatorname{Ext}^{g-1}_R(L,M/xM)\cong \operatorname{Ext}^g_R(L,M)\text{.}\)
Proof.
Let \(f:L\to M\) be a homomorphism. Then \(xf(L)=f(xL)=f(0)=0\text{.}\) As \(x\) is regular on \(M\text{,}\) we see that \(f(L)=0\text{.}\) Hence, \(\operatorname{Hom}_R(L,M)=0\) and so \(g>0\text{.}\) Now applying \(\operatorname{Hom}_R(L,-)\) to the s.e.s. \(0\to M\xrightarrow{x} M\to M/xM\to 0\) we get the long exact sequence
\begin{equation*}
\cdots\to \operatorname{Ext}^i_R(L,M)\xrightarrow{x} \operatorname{Ext}^i_R(L,M)\to \operatorname{Ext}^i_R(L,M/xM) \to \operatorname{Ext}^{i+1}_R(L,M)\xrightarrow{x} \cdots
\end{equation*}
Since \(xL=0\) we have \(x\operatorname{Ext}^i_R(L,M)=0\) for all \(i\) (cf. Grifo’s 915 notes, Exercise 73(c).) Thus, for each \(i\) we have an exact sequence
\begin{equation*}
0\to \operatorname{Ext}^i_R(L,M)\to \operatorname{Ext}^i_R(L,M/xM) \to \operatorname{Ext}^{i+1}_R(L,M)\to 0.
\end{equation*}
Both conclusions follow easily from this sequence.
Theorem 16.17.
Let \(R\) be a Noetherian ring, \(I\) an ideal of \(R\text{,}\) and \(M\) a finitely generated \(R\)-module such that \(IM\neq M\text{.}\) Let \(x_1,\dots,x_r\) be a maximal \(M\)-sequence contained in \(I\text{.}\) Then \(r=\sup\{n\mid \operatorname{Ext}^n_R(R/I,M)\neq 0\}.\) Consequently, all maximal \(M\)-sequences contained in \(I\) have the same length and
\begin{equation*}
\operatorname{grade}_IM=\inf\{n\mid \operatorname{Ext}^n_R(R/I,M)\neq 0\}.
\end{equation*}
In particular, \(\operatorname{grade}_I M<\infty\text{.}\)
Proof.
Notice that the right-hand-side is
\(g:=g(R/I,M)\text{.}\) Suppose
\(r=0\text{.}\) Since
\(IM\neq M\text{,}\) it must be that
\(I\) consists of zero-divisors on
\(M\text{.}\) Thus,
\(I\) is contained in the union of the associated primes of
\(M\) (Grifo 905, Theorem 6.27). By the prime avoidance lemma (Grifo 905, Theorem 3.29),
\(I\subseteq p\) for some associated prime
\(p\) of
\(M\text{.}\) Then the composition
\(R/I\to R/p\hookrightarrow M\) is nonzero. Hence,
\(\operatorname{Hom}_R(R/I,M)=\operatorname{Ext}^0_R(R/I,M)\neq 0\) and
\(g=0=r\text{.}\) Proceeding by induction on
\(r\text{,}\) we may assume
\(r>0\) and that the result holds for all finitely generated
\(R\)-modules
\(N\) with
\(IN\neq N\) and having a maximal
\(N\)-sequence of length at most
\(r-1\text{.}\) Let
\(N=M/x_1M\text{.}\) Then
\(IN\neq N\) and
\(x_2,\dots, x_r\) is a maximal
\(N\)-sequence contained in
\(I\text{.}\) Hence,
\(r-1=g(R/I,N)=g-1\) by Proposition
Proposition 16.16. Thus,
\(r=g\text{,}\) which completes the proof.
Corollary 16.18.
Let \(R\) be a Noetherian ring, \(I\) an ideal of \(R\text{,}\) and \(M\) a finitely generated \(R\)-module such that \(IM\neq M\text{.}\) If \(x\in I\) is a regular element on \(M\text{,}\) then
\begin{equation*}
\operatorname{grade}_{I/(x)}M/xM=\operatorname{grade}_I M/xM=\operatorname{grade}_I M-1.
\end{equation*}
Proof.
For the first equality, observe that any sequence
\({\bf y}\) in
\(I\) is an
\(M/xM\)-sequence if and only if its image
\({\bf \overline{y}}\) in
\(I/(x)\) is an
\(M/xM\)-sequence. For the second equality, we have by Theorem
Theorem 16.17 and Proposition
Proposition 16.16,
\begin{equation*}
\operatorname{grade}_I M/xM=g(R/I,M/xM)=g(R/I,M)-1=\operatorname{grade}_I M-1.
\end{equation*}
Definition 16.19.
Let \((R,m)\) be a local ring and \(M\) a finitely generated \(R\)-module. Then the depth of \(M\), denoted \(\operatorname{depth}M\text{,}\) is defined to be \(\operatorname{grade}_m M\text{;}\) i.e., the length of the longest \(M\)-sequence from \(R\text{.}\)
Note: By convention, the depth of the zero module is infinity, since \(\inf \emptyset = \infty\text{.}\)
Theorem 16.21. Ischebeck’s Theorem.
Let \((R,m)\) be a local ring and \(M\) and \(N\) finitely generated \(R\)-modules. Then \(\operatorname{Ext}^i_R(M,N)=0\) for \(i<\operatorname{depth}N - \operatorname{dim}M\text{.}\)
Proof.
First, if either module is zero the result trivially holds. So assume
\(M\) and
\(N\) are nonzero and let
\(d=\operatorname{dim}M\text{.}\) If
\(d=0\) then
\(M\) has finite length. Since
\(\operatorname{Ext}^i_R(R/m,N)=0\) for all
\(i<\operatorname{depth}N\) by Remark
Remark 16.20, we have
\(\operatorname{Ext}^i_R(M,N)=0\) for all
\(i<\operatorname{depth}N\) by Lemma
[cross-reference to target(s) "bass3" missing or not unique]
. Hence, the result holds for the case
\(d=0\text{.}\)
Suppose \(d>0\) and assume the result holds for all finitely generated modules of dimension less than \(d\text{.}\) Consider a filtration of \(M\text{:}\)
\begin{equation*}
0=M_0\subset M_1\subset \cdots \subset M_n=M
\end{equation*}
where \(M_j/M_{j-1}\cong R/p_j\) for some primes \(p_1,\dots,p_n\) of \(R\text{.}\) (See Grifo 905 :Theorem 6.33.) Note that each \(p_j\) contains \(\operatorname{Ann}_R M\text{,}\) so \(\operatorname{dim}R/p_j\leqslant d\) for all \(j\text{.}\) Hence, if we show \(\operatorname{Ext}^i_R(R/p_j,N)=0\) for all \(i<\operatorname{depth}N-\operatorname{dim}R/p_j\) and all \(j\text{,}\) then we’ll have \(\operatorname{Ext}^i_R(M_j/M_{j-1},N)=0\) for all \(i<\operatorname{depth}N-d\) and all \(j\text{.}\) Using the long exact sequences on \(\operatorname{Ext}\) arising from the short exact sequences
\begin{equation*}
0\to M_{j-1}\to M_j\to M_j/M_{j-1}\to 0,
\end{equation*}
we can conclude \(\operatorname{Ext}^i_R(M_j,N)=0\) for all \(i<\operatorname{depth}N-d\text{.}\) Since \(M_n=M\text{,}\) we’ll be done.
Thus, assume \(M\cong R/p\) for some prime \(p\) with \(\operatorname{dim}R/p=d>0\text{.}\) Since \(p\subsetneq m\text{,}\) let \(x\in m\setminus p\) and consider the short exact sequence
\begin{equation*}
0\to R/p\xrightarrow{x} R/p\to R/(p,x)\to 0.
\end{equation*}
Since \(\operatorname{dim}R/(p,x)\leqslant d-1\text{,}\) we have \(\operatorname{Ext}^i_R(R/(p,x), N)=0\) for all \(i<\operatorname{depth}N-d+1\) by the induction hypothesis. Rewriting this, we have \(\operatorname{Ext}^{i+1}_R(R/(p,x),N)=0\) for all \(i<\operatorname{depth}N-d\text{.}\) From the long exact sequence on \(\operatorname{Ext}\text{,}\) we have
\begin{equation*}
\operatorname{Ext}^i_R(R/p,N)\xrightarrow{x} \operatorname{Ext}^i_R(R/p,N)\to \operatorname{Ext}^{i+1}_R(R/(p,x),N).
\end{equation*}
By Nakayama’s Lemma, we see that \(\operatorname{Ext}^i_R(R/p,N)=0\) for all \(i<\operatorname{depth}N-d\text{.}\) This completes the proof.
Corollary 16.22.
Let \((R,m)\) be a local ring and \(M\) a nonzero finitely generated \(R\)-module. Then \(\operatorname{depth}M\leqslant\operatorname{dim}R/p\) for all \(p\in \operatorname{Ass}_R M\text{.}\) In particular, \(\operatorname{depth}M\leqslant\operatorname{dim}M\text{.}\)
Proof.
Let
\(p\in \operatorname{Ass}_R M\text{.}\) Then
\(\operatorname{Ext}^0_R(R/p,M)=\operatorname{Hom}_R(R/p,M)\neq 0\text{.}\) By Theorem
Theorem 16.21,
\(\operatorname{depth}M\leqslant\operatorname{dim}R/p\text{.}\) ◻
Example 16.23.
Let \(k\) be a field and \(R=k[x_1,\dots,x_{d+1}]/(x_1,\dots,x_{d+1})(x_{d+1})\text{.}\) Let \(m=(x_1,\dots,x_{d+1})R\text{.}\) Then \(R_m\) has depth zero and dimension \(d\text{.}\) To see \(\operatorname{depth}R_m=0\text{,}\) note that \(mx_{d+1}=0\) in \(R_m\text{,}\) so \(mR_m\) consists of zero-divisors. For \(\operatorname{dim}R_m\text{,}\) note that \(\operatorname{dim}k[x_1,\dots,x_{d+1}]_m\) is a \((d+1)\)-dimensional domain, so \(\operatorname{dim}R_m/I\leqslant d\) for any nonzero ideal \(I\text{.}\) On the other hand, \(R_m/(x_{d+1})R_m\cong k[x_1,\dots,x_d]_{(x_1,\dots,x_d)}\text{,}\) which has dimension \(d\text{.}\) Thus, \(\operatorname{dim}R_m\geqslant d\text{.}\)