In mathematics, the Koszul complex was first introduced to define a cohomology theory for Lie algebras, by Jean-Louis Koszul (see Lie algebra cohomology). It turned out to be a useful general construction in homological algebra. As a tool, its homology can be used to tell when a set of elements of a (local) ring is an M-regular sequence, and hence it can be used to prove basic facts about the depth of a module or ideal which is an algebraic notion of dimension that is related to but different from the geometric notion of Krull dimension. Moreover, in certain circumstances, the complex is the complex of syzygies, that is, it tells you the relations between generators of a module, the relations between these relations, and so forth.
Let A be a commutative ring and s: Ar → A an A-linear map. Its Koszul complex Ks is
wedgerAr \to wedger-1Ar \to … \to wedge1Ar \to wedge0Ar\simeqA
\alpha1\wedge … \wedge\alphak \mapsto \sum
k | |
i=1 |
(-1)i+1s(\alphai) \alpha1\wedge … \wedge\hat{\alpha}i\wedge … \wedge\alphak
\hat{ }
\wedge
Ar
Let M be a manifold, variety, scheme, ..., and A be the ring of functions on it, denoted
l{O}(M)
The map
s\colonAr\toA
f1,...,fr
l{O}(M) \stackrel{ ⋅ f}{\to} l{O}(M)
l{O}(M) \stackrel{ ⋅ (f1,...,f
r \to … \to l{O}(M) | |
r)}{\to} l{O}(M) |
r \stackrel{ ⋅ (f1,...,fr)}{\to} l{O}(M).
f1= … =fr=0
fi=0
In pictures: given functions
si
A/(s1,...,sr)
si=0
First, the Koszul complex Ks of (A,s) is a chain complex: the composition of any two maps is zero. Second, the map
Ks ⊗ Ks \to Ks (\alpha1\wedge … \wedge\alphak) ⊗ (\beta1\wedge … \wedge\beta\ell) \mapsto \alpha1\wedge … \wedge\alphak\wedge\beta1\wedge … \wedge\beta\ell
The Koszul complex is a tensor product: if
s=(s1,...,sr)
Ks
\simeq K | |
s1 |
⊗ … ⊗
K | |
sr |
⊗
When
s1,...,sr
Ks\toA/(s1,...,sr)
i(K | |
\operatorname{H} | |
s) = 0, |
i\ne0,
0(K | |
H | |
s) |
=A/(s1,...,sr)
The Koszul complex was first introduced to define a cohomology theory for Lie algebras, by Jean-Louis Koszul (see Lie algebra cohomology). It turned out to be a useful general construction in homological algebra. As a tool, its homology can be used to tell when a set of elements of a (local) ring is an M-regular sequence, and hence it can be used to prove basic facts about the depth of a module or ideal which is an algebraic notion of dimension that is related to but different from the geometric notion of Krull dimension. Moreover, in certain circumstances, the complex is the complex of syzygies, that is, it tells you the relations between generators of a module, the relations between these relations, and so forth.
Let R be a commutative ring and E a free module of finite rank r over R. We write
wedgeiE
s\colonE\toR
K\bullet(s)\colon0\towedgerE\overset{dr}\towedger-1E\to … \towedge1E\overset{d1}\toR\to0
dk
ei
dk(e1\wedge...\wedgeek)=
k | |
\sum | |
i=1 |
(-1)i+1s(ei)e1\wedge … \wedge\widehat{ei}\wedge … \wedgeek
\widehat{ ⋅ }
dk\circdk+1=0
Note that
wedge1E=E
d1=s
wedgerE\simeqR
If
E=Rr
s\colonRr\toR
s1,...,sr
K\bullet(s1,...,sr)=K\bullet(s).
If M is a finitely generated R-module, then one sets:
K\bullet(s,M)=K\bullet(s) ⊗ RM
(d ⊗ 1M)(v ⊗ m)=d(v) ⊗ m
The i-th homology of the Koszul complex
\operatorname{H}i(K\bullet(s,M))=\operatorname{ker}(di ⊗ 1M)/\operatorname{im}(di+1 ⊗ 1M)
E=Rr
s=[s1 … sr]
d1 ⊗ 1M
s\colonMr\toM,(m1,...,mr)\mapstos1m1+...+srmr
\operatorname{H}0(K\bullet(s,M))=M/(s1,...,sr)M=R/(s1,...,sr) ⊗ RM.
\operatorname{H}r(K\bullet(s,M))=\{m\inM:s1m=s2m=...=srm=0\}=\operatorname{Hom}R(R/(s1,...,sr),M).
Given a commutative ring R, an element x in R, and an R-module M, the multiplication by x yields a homomorphism of R-modules,
M\toM.
K(x,M)
H0(K(x,M))=M/xM,H1(K(x,M))=\operatorname{Ann}M(x)=\{m\inM,xm=0\},
K\bullet(x)
The Koszul complex for a pair
(x,y)\inR2
0\toR\xrightarrow{ d2 }R2\xrightarrow{ d1 }R\to0,
d1
d2
d1=\begin{bmatrix} x\\ y \end{bmatrix}
d2=\begin{bmatrix} -y&x \end{bmatrix}.
di
H1(K\bullet(x,y))
In the case that the elements
x1,x2,...,xn
If k is a field and
X1,X2,...,Xd
k[X1,X2,...,Xd]
K\bullet(Xi)
Xi
Let E be a finite-rank free module over R, let
s\colonE\toR
K(s,t)
(s,t)\colonE ⊕ R\toR
Using
wedgek(E ⊕ R)=
k | |
⊕ | |
i=0 |
wedgek-iE ⊗ wedgeiR=wedgekE ⊕ wedgek-1E
0\toK(s)\toK(s,t)\toK(s)[-1]\to0
[-1]
-1
dK(s)[-1]=-dK(s)
(x,y)
wedgekE ⊕ wedgek-1E
dK(s,((x,y))=(dK(s)x+ty,dK(s)[-1]y).
K(s,t)
t\colonK(s)\toK(s)
Taking the long exact sequence of homologies, we obtain:
… \to\operatorname{H}i(K(s))\overset{t}\to\operatorname{H}i(K(s))\to\operatorname{H}i(K(s,t))\to\operatorname{H}i-1(K(s))\overset{t}\to … .
\delta:\operatorname{H}i+1(K(s)[-1])=\operatorname{H}i(K(s))\to\operatorname{H}i(K(s))
\delta([x])=[dK(s,t)(y)]
K(s,t)
K(s,t)
dK(s,
\delta([x])=t[x]
The above exact sequence can be used to prove the following.
Proof by induction on r. If
r=1
\operatorname{H}1(K(x1;M))=\operatorname{Ann}M(x1)=0
\operatorname{H}i(K(x1,...,xr;M))=0
i\geq2
i=1
xr
\operatorname{H}0(K(x1,...,xr-1;M))=M/(x1,...,xr-1)M.
\square
Proof: By the theorem applied to S and S as an S-module, we see that
K(y1,...,yn)
S/(y1,...,yn)
K(y1,...,yn) ⊗ SM
K(y1,...,yn) ⊗ SM=K(x1,...,xn) ⊗ RM
\square
Proof: Let S = R[''y''<sub>1</sub>, ..., ''y''<sub>''n''</sub>]. Turn M into an S-module through the ring homomorphism S → R, yi → xi and R an S-module through . By the preceding corollary,
\operatorname{H}i(K(x1,...,xn) ⊗ M)=
S(R, | |
\operatorname{Tor} | |
i |
M)
\operatorname{Ann}S\left(\operatorname{Tor}
S(R, | |
i |
M)\right)\supset\operatorname{Ann}S(R)+\operatorname{Ann}S(M)\supset(y1,...,yn)+\operatorname{Ann}R(M)+(y1-x1,...,yn-xn).
\square
For a local ring, the converse of the theorem holds. More generally,
Proof: We only need to show 2. implies 1., the rest being clear. We argue by induction on r. The case r = 1 is already known. Let x denote x1, ..., xr-1. Consider
… \to\operatorname{H}1(K(x';M))\overset{xr}\to\operatorname{H}1(K(x';M))\to\operatorname{H}1(K(x1,...,xr;M))=0\toM/x'M\overset{xr}\to … .
xr
N=xrN
N=\operatorname{H}1(K(x';M))
N=0
xr
x1,...,xr
M/(x1,...,xr)M\ne0
\square
In general, if C, D are chain complexes, then their tensor product
C ⊗ D
(C ⊗ D)n=\sumiCi ⊗ Dj
dC(x ⊗ y)=dC(x) ⊗ y+(-1)|x|x ⊗ dD(y)
This construction applies in particular to Koszul complexes. Let E, F be finite-rank free modules, and let
s\colonE\toR
t\colonF\toR
K(s,t)
(s,t)\colonE ⊕ F\toR
K(s,t)\simeqK(s) ⊗ K(t).
-1
ds:\wedgeE\to\wedgeE
ds(x)=s(x)
|x|=1
ds(x\wedgey)=ds(x)\wedgey+(-1)|x|x\wedgeds(y)
ds\circds=0
ds
Now, we have
\wedge(E ⊕ F)=\wedgeE ⊗ \wedgeF
dK(s)(e ⊗ 1+1 ⊗ f)=dK(s)(e) ⊗ 1+1 ⊗ dK(t)(f)=s(e)+t(f)=dK(s,(e+f).
dK(s)
dK(s,
dK(s)=dK(s,.
Note, in particular,
K(x1,x2,...,xr)\simeqK(x1) ⊗ K(x2) ⊗ … ⊗ K(xr)
The next proposition shows how the Koszul complex of elements encodes some information about sequences in the ideal generated by them.
Proof: (Easy but omitted for now)
As an application, we can show the depth-sensitivity of a Koszul homology. Given a finitely generated module M over a ring R, by (one) definition, the depth of M with respect to an ideal I is the supremum of the lengths of all regular sequences of elements of I on M. It is denoted by
\operatorname{depth}(I,M)
M/(x1,...,xn)M
The Koszul homology gives a very useful characterization of a depth.
Proof: To lighten the notations, we write H(-) for H(K(-)). Let y1, ..., ys be a maximal M-regular sequence in the ideal I; we denote this sequence by
\underline{y}
l
\operatorname{H}i(\underline{y},x1,...,xl;M)
\operatorname{Ann}M/\underline{yM}(x1,...,xl)
i=l
i>l
l=0
\operatorname{H}l\left(\underline{y},x1,...,xl;M\right)=\operatorname{ker}\left(xl:\operatorname{Ann}M/\underline{yM}(x1,...,xl-1)\to\operatorname{Ann}M/\underline{yM}(x1,...,xl-1)\right)
\operatorname{Ann}M/\underline{yM}(x1,...,xl).
i>l
Now, it follows from the claim and the early proposition that
\operatorname{H}i(x1,...,xn;M)=0
\underline{y}
M/\underline{y}M
M/\underline{y}M
I\subsetak{p}=\operatorname{Ann}R(v)
0\nev\in\operatorname{Ann}M/\underline{yM}(I)\simeq\operatorname{H}n\left(x1,...,xn,\underline{y};M\right)=\operatorname{H}n-s(x1,...,xn;M) ⊗ \wedgesRs.
\square
There is an approach to a Koszul complex that uses a cochain complex instead of a chain complex. As it turns out, this results essentially in the same complex (the fact known as the self-duality of a Koszul complex).
Let E be a free module of finite rank r over a ring R. Then each element e of E gives rise to the exterior left-multiplication by e:
le:\wedgekE\to\wedgek+1E,x\mapstoe\wedgex.
e\wedgee=0
le\circle=0
0\toR\overset{1\mapstoe}\to\wedge1E\overset{le}\to\wedge2E\to … \to\wedgerE\to0
0\to(\wedgerE)*\to(\wedger-1E)*\to … \to(\wedge2E)*\to(\wedge1E)*\toR\to0
\wedgekE\simeq(\wedger-kE)*\simeq\wedger-k(E*)
(\wedgeE,le)
The Koszul complex is essential in defining the joint spectrum of a tuple of commuting bounded linear operators in a Banach space.
(x,y)=(e1+\epsilon)\wedgee2\wedge … \wedgeek\in\wedgek(E ⊕ R)
R\simeqR\epsilon\subsetE ⊕ R
dK(s,((x,y))=(s(e1)+t)e2\wedge … \wedgeek+
k | |
\sum | |
i=2 |
(-1)is(ei)(e1+\epsilon)\wedgee2\wedge … \widehat{ei} … \wedgeek
(dK(s)x+ty,-dK(s)y)