Datasets:
Tasks:
Text Generation
Modalities:
Text
Sub-tasks:
language-modeling
Languages:
English
Size:
100K - 1M
License:
\chapter{Homological Algebra} | |
\label{homological} | |
Homological algebra begins with the notion of a \emph{differential object,} | |
that is, an object with an endomorphism $A \stackrel{d}{\to} A$ such that $d^2 = | |
0$. This equation leads to the obvious inclusion $\im(d) \subset \ker(d)$, but | |
the inclusion generally is not equality. We will find that the difference | |
between $\ker(d)$ and $\im(d)$, called the \emph{homology}, is a highly useful | |
variant of a differential object: its first basic property is that if an exact | |
sequence | |
\[ 0 \to A' \to A \to A'' \to 0 \] | |
of differential objects is given, the homology of $A$ is related to that of | |
$A', A''$ through a long exact sequence. The basic example, and the one we | |
shall focus on, is where $A$ is a | |
chain complex, and $d$ the usual differential. | |
In this case, homology simply measures the failure of a complex to be exact. | |
After introducing these preliminaries, we develop the theory of \emph{derived | |
functors}. Given a functor that is only left or right-exact, derived functors | |
allow for an extension of a partially exact sequence to a long exact sequence. | |
The most important examples to us, $\mathrm{Tor}$ and $\mathrm{Ext}$, provide | |
characterizations of flatness, projectivity, and injectivity. | |
\section{Complexes} | |
\subsection{Chain complexes} | |
The chain complex is the most fundamental construction in | |
homological algebra. | |
\begin{definition} Let $R$ be a ring. A \textbf{chain complex} is a collection | |
of $R$-modules | |
$\{C_i\}$ (for $i \in \mathbb{Z}$) | |
together with boundary | |
operators | |
$\partial_i:C_i\rightarrow C_{i-1}$ such that | |
$\partial_{i-1}\partial_i=0$. The boundary map is also | |
called the | |
\textbf{differential.} Often, notation is abused and the indices for | |
the boundary map are dropped. | |
A chain complex is often simply denoted $C_*$. | |
\end{definition} | |
In practice, one often has that $C_i = 0$ for $i<0$. | |
\begin{example} All exact sequences are chain complexes. | |
\end{example} | |
\begin{example} Any sequence of abelian groups $\left\{C_i\right\}_{i \in | |
\mathbb{Z}}$ with the boundary operators | |
identically zero forms a chain complex. | |
\end{example} | |
We will see plenty of more examples in due time. | |
At each stage, elements in the image of the boundary $C_{i+1} \to C_i$ lie in | |
the kernel of $\partial_i: C_i \to C_{i-1}$. Let us recall that a chain | |
complex is \emph{exact} if the kernel and the image coincide. In general, a | |
chain complex need not be exact, and this failure of exactness is measured by | |
its homology. | |
\begin{definition} | |
Let $C_*$ | |
The submodule of cycles $Z_i\subset C_i$ is | |
the kernel $\ker(\partial_i)$. The submodule of boundaries | |
$B_i\subset C_i$ is the image $Im(\partial_{i+1})$. Thus | |
homology is said to be ``cycles mod boundaries,'' i.e. | |
$Z_i/B_i$. | |
\end{definition} | |
To further simplify notation, often all differentials regardless | |
of what chain complex they are part of are denoted $\partial$, | |
thus the commutativity relation on chain maps is | |
$f\partial=\partial f$ with indices and distinction between the | |
boundary operators dropped. | |
\begin{definition} Let $C_*$ be a chain complex with boundary | |
map $\partial_i$. | |
We define the \textbf{homology} of the complex $C_*$ via | |
$H_i(C_*)=\ker(\partial_i)/Im(\partial_{i+1})$. | |
\end{definition} | |
\begin{example} In a chain complex $C_*$ where all the boundary | |
maps are | |
trivial, $H_i(C_*)=C_i$. | |
\end{example} | |
Often we will bundle all the modules $C_i$ of a chain complex | |
together to form a graded module $C_*=\bigoplus_i C_i$. In this | |
case, the boundary operator is a | |
endomorphism that takes elements from degree $i$ to degree | |
$i-1$. Similarly, we | |
often bundle together all the homology modules to give a graded | |
homology module | |
$H_*(C_*)=\bigoplus_i H_i(C_*)$. | |
\begin{definition} | |
A \textbf{differential module} is a module $M$ together with a morphism $d: | |
M\to M$ such that $d^2 =0$. | |
\end{definition} | |
Thus, given a chain complex $C_*$, the module $\bigoplus C_i$ is a | |
differential module with the direct sum of all the differentials $\partial_i$. | |
A chain complex is just a special kind of differential module, one where the | |
objects are graded and the differential drops the grading by one. | |
\subsection{Functoriality} | |
We have defined chain complexes now, but we have no notion of a morphism | |
between chain complexes. | |
We do this next; it turns out that chain complexes form a category when morphisms | |
are appropriately defined. | |
\begin{definition} A \textbf{morphism} of chain complexes $f:C_*\rightarrow | |
D_*$, or a \textbf{chain map}, is a sequence of maps $f_i:C_i\rightarrow | |
D_i$ such that $f\partial = \partial' f$ where $\partial$ is the | |
boundary map of $C_*$ and $\partial'$ of $D_*$ (again we are | |
abusing notation and dropping indices). | |
\end{definition} | |
There is thus a \emph{category} of chain complexes where the morphisms are | |
chain maps. | |
One can make a similar definition for differential modules. If $(M, d)$ and | |
$(N,d')$ are differential modules, then a \emph{morphism of differential | |
modules} $(M,d) \to (N,d')$ is a morphism of modules $M \to N$ such that the diagram | |
\[ | |
\xymatrix{ | |
M \ar[d] \ar[r]^d & M \ar[d] \\ | |
N \ar[r]^{d'} & N | |
} | |
\] | |
commutes. | |
There is therefore a category of differential modules, and the map $C_* \to | |
\bigoplus C_i$ gives a functor from the category of chain complexes to that of | |
differential modules. | |
\begin{proposition} A chain map $C_* \to D_*$ induces a map in homology $H_i(C) | |
\to H_i(D)$ for each $i$; thus homology is a covariant functor from | |
the category of chain complexes to the category of graded | |
modules. | |
\end{proposition} | |
More precisely, each $H_i$ is a functor from chain complexes to modules. | |
\begin{proof} | |
Let $f:C_*\rightarrow D_*$ be a chain map. Let $\partial$ and | |
$\partial'$ be the differentials for $C_*$ and $D_*$ | |
respectively. Then we have a commutative diagram: | |
\begin{equation} | |
\begin{CD} | |
C_{i+1} @>\partial_{i+1}>> C_i @>>\partial_i> C_{i-1}\\ | |
@VV f_{i+1} V @VV f_i V @VVf_{i-1} V\\ | |
D_{i+1} @>\partial'_{i+1}>> D_i @>>\partial'_i> D_{i-1} | |
\end{CD} | |
\end{equation} | |
Now, in order to check that a chain map $f$ induces a map $f_*$ | |
on homology, we need to check that $f_*(Im(\partial))\subset | |
Im(\partial')$ and $f_*(\ker(\partial))\subset | |
\ker(\partial)$. We first check the condition on images: we want | |
to look at $f_i(Im(\partial_{i+1}))$. By commutativity of $f$ | |
and the boundary maps, this is equal to | |
$\partial'_{i+1}(Im(f_{i+1})$. Hence we have | |
$f_i(Im(\partial_{i+1}))\subset Im(\partial_{i+1}')$. For the | |
condition on kernels, let $x\in \ker(\partial_i)$. Then by | |
commutativity, $\partial'_i(f_i(x))=f_{i-1}\partial_i(x)=0$. | |
Thus we have that $f$ induces for each $i$ a homomorphism | |
$f_i:H_i(C_*)\rightarrow H_i(D_*)$ and hence it induces a | |
homomorphism on homology as a graded module. \end{proof} | |
\begin{exercise} | |
Define the \emph{homology} $H(M)$ of a differential module $(M, d)$ via $\ker d / \im | |
d$. Show that $M \mapsto H(M)$ is a functor from differential modules to | |
modules. | |
\end{exercise} | |
\subsection{Long exact sequences} | |
\add{OMG! We have all this and not the most basic theorem of them all.} | |
\begin{definition} If $M$ is a complex then for any integer $k$, we define a new complex $M[k]$ by shifting indices, i.e. $(M[k])^i:=M^{i+k}$.\end{definition} | |
\begin{definition} If $f:M\rightarrow N$ is a map of complexes, we define a complex $\mathrm{Cone}(f):=\{N^i\oplus M^{i+1}\}$ with differential | |
$$d(n^i,m^{i+1}):= (d_N^i(n_i)+(-1)^i\cdot f(m^{i+1}, d_M^{i+1}(m^{i+1}))$$ | |
\end{definition} | |
Remark: This is a special case of the total complex construction to be seen later. | |
\begin{proposition} A map $f:M\rightarrow N$ is a quasi-isomorphism if and only if $\mathrm{Cone}(f)$ is acyclic.\end{proposition} | |
\begin{proposition} Note that by definition we have a short exact sequence of complexes | |
$$0\rightarrow N\rightarrow \mathrm{Cone}(f)\rightarrow M[1]\rightarrow 0$$ | |
so by Proposition 2.1, we have a long exact sequence | |
$$\dots \rightarrow H^{i-1}(\mathrm{Cone}(f))\rightarrow H^{i}(M)\rightarrow H^{i}(N)\rightarrow H^{i}(\mathrm{Cone}(f))\rightarrow\dots$$ | |
so by exactness, we see that $H^i(M)\simeq H^i(N)$ if and only if $H^{i-1}(\mathrm{Cone}(f))=0$ and $H^i(\mathrm{Cone}(f))=0$. Since this is the case for all $i$, the claim follows. $\blacksquare$ | |
\end{proposition} | |
\subsection{Cochain complexes} | |
Cochain complexes are much like chain complexes except the | |
arrows point in the | |
opposite direction. | |
\begin{definition} A \textbf{cochain complex} is a sequence of modules | |
$C^i$ for $i \in \mathbb{Z}$ with \textbf{coboundary operators}, also | |
called | |
\textbf{differentials}, $\partial^i:C^i\rightarrow C^{i+1}$ such that | |
$\partial^{i+1}\partial^i=0$. \end{definition} | |
The theory of cochain complexes is entirely dual to that of chain complexes, | |
and we shall not spell it out in detail. | |
For instance, we can form a category of cochain complexes and | |
\textbf{chain maps} (families of morphisms commuting with the | |
differential). Moreover, given a cochain complex $C^*$, we | |
define the | |
\textbf{cohomology objects} to be | |
$h^i(C^*)=\ker(\partial^i)/Im(\partial^{i-1})$; one obtains cohomology | |
functors. | |
It should be noted that the long exact sequence in cohomology runs in the | |
opposite direction. | |
If $0 \to C_*' \to C_* \to C_*'' \to 0$ is a short exact sequence of cochain | |
complexes, we get a long exact sequence | |
\[ \dots \to H^i(C' ) \to H^i(C) \to H^{i}(C'') \to H^{i+1}(C' ) \to H^{i+1}(C) \to | |
\dots. \] | |
Similarly, we can also turn cochain complexes and cohomology | |
modules into a | |
graded module. | |
Let us now give a standard example of a cochain complex. | |
\begin{example}[The de Rham complex] Readers unfamiliar with differential | |
forms may omit this example. Let $M$ be a smooth manifold. For each $p$, let | |
$C^p(M)$ be the $\mathbb{R}$-vector space of smooth $p$-forms on $M$. | |
We can make the $\left\{C^p(M)\right\}$ into a complex by defining the maps | |
\[ C^p(M) \to C^{p+1}(M) \] | |
via $\omega \to d \omega$, for $d$ the exterior derivative. | |
(Note that $d^2 = 0$.) This complex is called the \textbf{de Rham complex} of | |
$M$, and its cohomology is called the \textbf{de Rham cohomology.} It is known | |
that the de Rham cohomology is isomorphic to singular cohomology with real | |
coefficients. | |
It is a theorem, which we do not prove, that the de Rham cohomology is | |
isomorphic to the singular cohomology of $M$ with coefficients in $\mathbb{R}$. | |
\end{example} | |
\subsection{Long exact sequence} | |
\subsection{Chain Homotopies} | |
In general, two maps of complexes $C_* \rightrightarrows D_*$ need not be | |
equal to induce the same morphisms in homology. It is thus of interest to | |
determine conditions when they do. One important condition is given by chain | |
homotopy: chain homotopic maps are indistinguishable in homology. In algebraic | |
topology, this fact is used to show that singular homology is a homotopy | |
invariant. | |
We will find it useful in showing that the construction (to be given later) of a | |
projective resolution is essentially unique. | |
\begin{definition} Let $C_*, D_*$ be chain complexes with differentials $d_i$. A chain homotopy between two chain maps | |
$f,g:C_*\rightarrow D_*$ is a series of homomorphisms | |
$h^i:C^i\rightarrow D^{i-1}$ satisfying $f^i-g^i=d h^i+ | |
h^{n+1}d$. Again often notation is abused and the | |
condition is written $f-g=d h + | |
hd$. | |
\end{definition} | |
\begin{proposition} If two morphisms of complexes $f,g: C_* \to D_*$ are chain homotopic, they are taken | |
to the same induced map after applying the homology functor. | |
\end{proposition} | |
\begin{proof} | |
Write $\left\{d_i\right\}$ for the various differentials (in both complexes). | |
Let $m\in Z_i(C)$, the group of $i$-cycles. | |
Suppose there is a chain homotopy $h$ between $f,g$ (that is, a set of | |
morphisms $C_i \to D_{i-1}$). | |
Then | |
$$f^i(m)-g^i(m)= h^{i+1}\circ d^i(m) + d^{i-1}\circ h^i(m)= d^{i-1}\circ H^i(m) \in \Im(d^{i-1})$$ | |
which is zero in the cohomology $H^i(D)$. | |
\end{proof} | |
\begin{corollary} If two chain complexes are chain homotopically equivalent | |
(there are maps $f: C_*\rightarrow D_*$ and $g:D_*\rightarrow | |
C_*$ such that both $fg$ and $gf$ are chain homotopic to the | |
identity), they have isomorphic homology. | |
\end{corollary} | |
\begin{proof} | |
Clear. | |
\end{proof} | |
\begin{example} Not every quasi-isomorphism is a homotopy equivalence. Consider the complex | |
$$\dots \rightarrow 0\rightarrow\mathbb{Z}/{\cdot 2}\rightarrow \mathbb{Z}\rightarrow 0\rightarrow 0\rightarrow\dots$$ | |
so $H^0=\mathbb{Z}/2\mathbb{Z}$ and all cohomologies are 0. We have a quasi-isomorphism from the above complex to the complex | |
$$\dots \rightarrow 0\rightarrow 0 \rightarrow \mathbb{Z}/2\mathbb{Z}\rightarrow 0\rightarrow 0\rightarrow\dots$$ | |
but no inverse can be defined (no map from $\mathbb{Z}/2\mathbb{Z}\rightarrow \mathbb{Z}$). | |
\end{example} | |
\begin{proposition} Additive functors preserve chain homotopies | |
\end{proposition} | |
\begin{proof} Since an additive functor $F$ is a homomorphism on $Hom(-,-)$, | |
the chain homotopy condition will be preserved; in | |
particular, if $t$ is a chain homotopy, then $F(t)$ is a chain | |
homotopy. | |
\end{proof} | |
In more sophisticated homological theory, one often makes the | |
definition of the ``homotopy category of chain complexes.'' | |
\begin{definition} The homotopy category of chain complexes is | |
the category $hKom(R)$ where objects are chain complexes of | |
$R$-modules and morphisms are chain maps modulo chain homotopy. | |
\end{definition} | |
\subsection{Topological remarks} | |
\add{add more detail} | |
The first homology theory | |
to be developed was simplicial homology - the study of homology | |
of simplicial | |
complexes. To be simple, we will not develop the general theory | |
and instead | |
motivate our definitions with a few basic examples. | |
\begin{example} Suppose | |
our simplicial complex has one line segment with both ends | |
identified at | |
one point $p$. Call the line segment $a$. The $n$-th homology | |
group of this | |
space roughly counts how many ``different ways'' there are of | |
finding $n$ | |
dimensional sub-simplices that have no boundary that aren't the | |
boundary of | |
any $n+1$ dimensional simplex. For the circle, notice that for | |
each integer, | |
we can find such a way (namely the simplex that wraps counter | |
clockwise that | |
integer number of times). The way we compute this is we look at | |
the free abelian group generated by $0$ simplices, and $1$ | |
simplices (there are no simplices of | |
dimension $2$ or higher so we can ignore that). We call these | |
groups $C_0$ and | |
$C_1$ respectively. There is a boundary map $\partial_1: | |
C_1\rightarrow C_0$. | |
This boundary map takes a $1$-simplex and associates to it its | |
end vertex minus | |
its starting vertex (considered as an element in the free | |
abelian group on | |
vertices of our simplex). In the case of the circle, since there | |
is only one | |
$1$-simplex and one $0$-simplex, this map is trivial. We then | |
get our homology | |
group by looking at $\ker(\partial_1)$. In the case that there | |
is a nontrivial | |
boundary map $\partial_2: C_2\rightarrow C_1$ (which can only | |
happen when our | |
simplex is at least $2$-dimensional), we have to take the | |
quotient | |
$\ker(\partial_1)/\ker(\partial_2)$. This motivates us to define | |
homology in a | |
general setting. | |
\end{example} | |
Originally homology was | |
intended to be a homotopy invariant meaning that space with the | |
same homotopy type would have isomorphic homology modules. In fact, any | |
homotopy induces what is now known as a chain homotopy on the simplicial chain | |
complexes. | |
\begin{exercise}[Singular homology] Let $X$ be a topological | |
space and let $S^n$ be the set of all continuous maps | |
$\Delta^n\rightarrow X$ where $\Delta^n$ is the convex hull of | |
$n$ distinct points and the origin with orientation given by an | |
ordering of the $n$ vertices. Define $C_n$ to be the free | |
abelian group generated by elements of $S^n$. Define | |
$\Delta^n_{\hat{i}}$ to be the face of $\Delta^n$ obtained by | |
omitting the $i$-th vertex from the simplex. We can then | |
construct a boundary map $\partial_n:C_n\rightarrow C_{n-1}$ to | |
take a map $\sigma^n:\Delta^n\rightarrow X$ to | |
$\sum_{i=0}^n(-1)^i\sigma^n|_{\Delta^n_{\hat{i}}}$. Verify that | |
$\partial^2=0$ (hence making $C_*$ into a chain complex known as | |
the ``singular chain complex of $X$''. Its homology groups are | |
the ``singular homology groups''. \end{exercise} | |
\begin{exercise} Compute the singular homology groups of a | |
point. \end{exercise} | |
\section{Derived functors} | |
\subsection{Projective resolutions} | |
Fix a ring $R$. | |
Let us recall (\rref{projectives}) that an $R$-module $P$ is called | |
\emph{projective} if the functor $N \to \hom_R(P,N)$ (which is always | |
left-exact) is exact. | |
Projective objects are useful in defining chain exact sequences | |
known as ``projective resolutions.'' In the theory of derived functors, the | |
projective resolution of a module $M$ is in some sense a replacement for $M$: | |
thus, we want it to satisfy some uniqueness and existence properties. The | |
uniqueness is not quite true, but it is true modulo chain equivalence. | |
\begin{definition} Let $M$ be an arbitrary module, a projective | |
resolution of | |
$M$ is an exact sequence | |
\begin{equation} \cdots\rightarrow P_i\rightarrow | |
P_{i-1}\rightarrow | |
P_{i-2}\cdots\rightarrow P_1\rightarrow P_0\rightarrow M | |
\end{equation} where | |
the $P_i$ are projective modules. \end{definition} | |
\begin{proposition} Any module admits a projective resolution. \end{proposition} | |
The proof will even show that we can take a \emph{free} resolution. | |
\begin{proof} | |
We construct the resolution inductively. | |
First, we take a projective module $P_0$ with $P_0 \twoheadrightarrow N$ | |
surjective by the previous part. Given a portion of the resolution | |
\[ P_n \to P_{n-1} \to \dots \to P_0 \twoheadrightarrow N \to 0 \] | |
for $n \geq 0$, which is exact at each step, we consider $K = \ker(P_n \to | |
P_{n-1})$. The sequence | |
\[ 0 \to K \to P_n \to P_{n-1} \to \dots \to P_0 \twoheadrightarrow N \to 0 \] | |
is exact. So if $P_{n+1}$ is chosen such that it is projective and there is an | |
epimorphism | |
\( P_{n+1} \twoheadrightarrow K, \) | |
(which we can construct by \rref{freesurjection}), then | |
\[ P_{n+1} \to P_n \to \dots \] | |
is exact at every new step by construction. We can repeat this inductively and | |
get a full projective resolution. | |
\end{proof} | |
Here is a useful observation: | |
\begin{proposition} | |
If $R$ is noetherian, and $M$ is finitely generated, then we can | |
choose a | |
projective resolution where each $P_i$ is finitely generated. | |
\end{proposition} | |
We can even take a resolution consisting of finitely generated free modules. | |
\begin{proof} | |
To say that $M$ is finitely generated is to say that it is a | |
quotient of a free module on | |
finitely many generators, so we can take $P_0$ free and finitely generated. The kernel | |
of $P_0 \to M$ | |
is finitely generated by noetherianness, and we can proceed as | |
before, at each step | |
choosing a finitely generated object. | |
\end{proof} | |
\begin{example} The abelian group $\mathbb{Z}/2$ has the free | |
resolution $0\rightarrow\cdots | |
0\rightarrow\mathbb{Z}\rightarrow\mathbb{Z}\rightarrow\mathbb{Z}/2$. | |
Similarly, since any finitely generated abelian group can be | |
decomposed into the direct sum of torsion subgroups and free | |
subgroups, all finitely generated abelian groups admit | |
a free resolution of length two. | |
Actually, over a principal ideal domain $R$ (e.g. $R=\mathbb{Z}$), | |
\emph{every} module admits a free resolution of length two. The reason is that | |
if $F \twoheadrightarrow M$ is a surjection with $F$ free, then the kernel $F' | |
\subset F$ is free by a general fact (\add{citation needed}) that a submodule | |
of a free module is free (if one works over a PID). So we get a free | |
resolution of the type | |
\[ 0 \to F' \to F \to M \to 0. \] | |
\end{example} | |
In general, projective resolutions are not at all unique. | |
Nonetheless, they \emph{are} unique up to chain homotopy. Thus a projective | |
resolution is a rather good ``replacement'' for the initial module. | |
\begin{proposition} | |
Let $M, N$ be modules and let $P_* \to M, P'_* \to N$ be projective | |
resolutions. Let $f: M \to N$ be a morphism. Then there is a morphism | |
\[ P_* \to P'_* \] | |
such that the following diagram commutes: | |
\[ | |
\xymatrix{ | |
\dots \ar[r] & P_1 \ar[r] \ar[d] & P_0 \ar[r] \ar[d] & M \ar[d]^f \\ | |
\dots \ar[r] & P'_1 \ar[r] & P'_0 \ar[r] & N | |
} | |
\] | |
This morphism is unique up to chain homotopy. | |
\end{proposition} | |
\begin{proof} | |
Let $P_* \to M$ and $P'_* \to N$ be projective resolutions. We will define a | |
morphism of complexes $P_* \to P'_* $ such that the diagram commutes. | |
Let the boundary maps in $P_*, P'_*$ be denoted $d$ (by abuse of notation). | |
We have an exact diagram | |
\[ | |
\xymatrix{ | |
\dots \ar[r] & P_n \ar[r]^d & P_{n-1} \ar[r]^d & \dots \ar[r]^d & P_0 | |
\ar[r]& M \ar[d]^{f} \ar[r] & 0 \\ | |
\dots \ar[r] & P'_n \ar[r]^d & P'_{n-1} \ar[r] & \dots \ar[r]^d & P'_0 \ar[r] & N \ar[r] & 0 | |
} | |
\] | |
Since $P'_0 \twoheadrightarrow N$ is an epimorphism, the map $P_0 \to M \to N$ lifts | |
to a map $P_0 \to P'_0$ making the diagram | |
\[ \xymatrix{ | |
P_0 \ar[d] \ar[r] & M \ar[d]^{f} \\ | |
P'_0 \ar[r] & N | |
}\] | |
commute. | |
Suppose we have defined maps $P_i \to P'_i$ for $i \leq n$ such that the | |
following diagram commutes: | |
\[ | |
\xymatrix{ | |
P_n \ar[r]^d \ar[d] & P_{n-1} \ar[r]^d \ar[d] & \dots \ar[r]^d & P_0 | |
\ar[d] \ar[r]& M \ar[d]^{f} \ar[r] & 0 \\ | |
P'_n \ar[r]^d & P'_{n-1} \ar[r] & \dots \ar[r]^d & P'_0 \ar[r] & N \ar[r] & 0 | |
} | |
\] | |
Then we will define $P_{n+1} \to P'_{n+1}$, after which induction will prove | |
the existence of a map. To do this, note that | |
the map | |
\[ P_{n+1} \to P_n \to P'_n \to P'_{n-1} \] | |
is zero, because this is the same as $P_{n+1} \to P_n \to P_{n-1} \to P'_{n-1}$ | |
(by induction, the diagrams before $n$ commute), and this is zero because two | |
$P$-differentials were composed one after another. In particular, in the diagram | |
\[ | |
\xymatrix{ | |
P_{n+1} \ar[r] & P_n \ar[d] \\ | |
P'_{n+1} \ar[r] & P'_n | |
}, | |
\] | |
the image in $P'_n$ of $P_{n+1}$ lies in the kernel of $P'_n \to P'_{n-1}$, | |
i.e. in the image $I$ of $P'_{n+1}$. The exact diagram | |
\[ | |
\xymatrix{ | |
& P_{n+1} \ar[d] \\ | |
P'_{n+1} \ar[r] & I \ar[r] & 0 | |
} | |
\] | |
shows that we can lift $P_{n+1} \to I$ to $P_{n+1} \to P'_{n+1}$ (by | |
projectivity). This implies that we can continue the diagram further and get a | |
morphism $P_* \to P'_* $ of complexes. | |
Suppose $f, g: P_* \to P'_*$ are two morphisms of the projective resolutions | |
making $$\xymatrix{ | |
P_0 \ar[r] \ar[d] & M \ar[d] \\ | |
P'_0 \ar[r] & N | |
}$$ commute. We will show that $f,g$ are chain homotopic. | |
For this, | |
we start by defining $D_0: P_0 \to P'_1$ such that $dD_0 = f-g: P_0 \to P'_0$. | |
This we can do because $f-g$ sends $P_0$ into $\ker(P'_0 \to N)$, i.e. into the | |
image of $P'_1 \to P'_0$, and $P_0$ is projective. | |
Suppose we have defined chain-homotopies $D_i: P_{i} \to P_{i+1}$ for $i \leq | |
n$ such that $dD_i + D_{i-1}d = f-g$ for $i \leq n$. We will define $D_{n+1}$. | |
There is a diagram | |
\[ | |
\xymatrix{ | |
& P_{n+1} \ar[d] \ar[r] & P_n \ar[ld]^{D_n}\ar[d] \ar[r] & P_{n-1} | |
\ar[ld]^{D_{n-1}} \ar[d] \\ | |
P'_{n+2} \ar[r] & P'_{n+1} \ar[r] & P'_n \ar[r] & P'_{n-1} \\ | |
}\] | |
where the squares commute regardless of whether you take the vertical maps to | |
be $f$ or $g$ (provided that the choice is consistent). | |
We would like to define $D_{n+1}: P_n \to P'_{n+1}$. | |
The key condition we need satisfied is that | |
\[ d D_{n+1} = f - g - D_n d. \] | |
However, we know that, by the inductive hypothesis on the $D$'s | |
\[ d( f- g - D_{n}d) = fd - gd - dD_n d = fd - gd - (f-g)d + D_n dd = 0. \] | |
In particular, $f-g - D_n d$ lies in the image of $P'_{n+1} \to P'_n$. | |
The projectivity of $P_n$ ensures that we can define $D_{n+1}$ satisfying the | |
necessary condition. | |
\end{proof} | |
\begin{corollary} | |
Let $P_* \to M, P'_* \to M$ be projective resolutions of $M$. Then there are | |
maps $P_* \to P'_*, P'_* \to P_* $ under $M$ such that the compositions are | |
chain homotopic to the identity. | |
\end{corollary} | |
\begin{proof} | |
Immediate. | |
\end{proof} | |
\subsection{Injective resolutions} | |
One can dualize all this to injective resolutions. \add{do this} | |
\subsection{Definition} | |
Often in homological algebra, we see that ``short exact | |
sequences induce long exact sequences.'' Using the theory of | |
derived functors, we can make this formal. | |
Let us work in the category of modules over a ring $R$. Fix two such categories. | |
Recall that a right-exact functor $F$ (from the category of modules over a | |
ring to the category of modules over another ring) is an additive functor | |
such that for every short | |
exact sequence $0\rightarrow A\rightarrow B\rightarrow | |
C\rightarrow 0$, we get a exact sequence $F(A)\rightarrow | |
F(B)\rightarrow F(C)\rightarrow 0$. | |
We want a natural way to continue this exact sequence to the | |
left; one way of doing this is to define the left derived | |
functors. | |
\begin{definition} Let $F$ be a right-exact functor and | |
$P_*\rightarrow M$ are projective resolution. We can form a | |
chain complex $F(P_*)$ whose object in degree $i$ is $F(P_i)$ | |
with boundary maps $F(\partial)$. The homology of this chain | |
complex denoted $L_iF$ are the left derived functors. | |
\end{definition} | |
For this definition to be useful, it is important to verify that | |
deriving a functor yields functors independent on choice of | |
resolution. This is clear by \rref{}. | |
\begin{theorem} The following properties characterize derived | |
functors: \begin{enumerate} | |
\item{ $L_0F(-)=F(-)$ } | |
\item{ Suppose $0\rightarrow A\rightarrow B\rightarrow | |
C\rightarrow 0$ is an exact sequence and $F$ a right-exact | |
functor; the left derived functors fit into the following exact | |
sequence: | |
\begin{equation} \cdots L_iF(A)\rightarrow L_iF(B)\rightarrow | |
L_iF(C)\rightarrow L_{i-1}F(A)\cdots\rightarrow | |
L_1(C)\rightarrow L_0F(A)\rightarrow L_0F(B)\rightarrow | |
L_0F(C)\rightarrow 0 \end{equation}} | |
\end{enumerate} | |
\end{theorem} | |
\begin{proof} The second property is the hardest to prove, but | |
it is by far the most useful; it is essentially an application | |
of the snake lemma. \end{proof} | |
One can define right derived functors analogously; if one has a | |
left exact functor (an additive functor that takes an exact | |
sequence $0\rightarrow A\rightarrow B\rightarrow C\rightarrow 0$ to | |
$0\rightarrow F(A)\rightarrow F(B)\rightarrow F(C)$), we can | |
pick an injective resolution instead (the injective criterion is simply the | |
projective criterion with arrows reversed). If | |
$M\rightarrow I^*$ is a injective resolution then the cohomology of the chain | |
complex $F(I^*)$ gives the right derived functors. | |
However, variance must also be taken into consideration so the | |
choice of whether or not to use a projective or injective | |
resolution is of importance (in all of the above, functors were | |
assumed to be covariant). In the following, we see an example of when right | |
derived functors can be computed using projective | |
resolutions. | |
\subsection{$\ext$ functors} | |
\begin{definition} The right derived functors of $Hom(-,N)$ are | |
called the $Ext$-modules denoted $Ext^i_R(-,N)$. | |
\end{definition} | |
We now look at the specific construction: | |
Let $M, M'$ be $R$-modules. Choose a projective resolution | |
\[ \dots \to P_2 \to P_1 \to P_0 \to M \to 0 \] | |
and consider what happens when you hom this resolution into $N$. | |
Namely, we can | |
consider $\hom_R(M,N)$, which is the kernel of $\hom(P_0, M) | |
\to\hom(P_1, M) $ | |
by exactness of the sequence | |
\[ 0 \to \hom_R(M,N) \to \hom_R(P_0, N) \to \hom_R(P_1, N) . \] | |
You might try to continue this with the sequence | |
\[ 0 \to \hom_R(M,N) \to \hom_R(P_0, N) \to \hom_R(P_1, N) \to | |
\hom_R(P_2, N) | |
\to \dots. \] | |
In general, it won't be exact, because $\hom_R$ is only | |
left-exact. But it is a | |
chain complex. You can thus consider the homologies. | |
\begin{definition} | |
The homology of the complex $\{\hom_R(P_i, N)\}$ is denoted | |
$\ext^i_R(M,N)$. By | |
definition, this is $\ker(\hom(P_i,N) \to \hom(P_{i+1}, | |
N))/\im(\hom(P_{i-1}, | |
N) \to \hom(P_i,N))$. This is an $R$-module, and is called the | |
$i$th ext group. | |
\end{definition} | |
Let us list some properties (some of these properties are just | |
case-specific examples of general properties of derived | |
functors) | |
\begin{proposition} | |
$\ext_R^0(M,N) = \hom_R(M,N)$. | |
\end{proposition} | |
\begin{proof} | |
This is obvious from the left-exactness of $\hom(-,N)$. (We | |
discussed this.) | |
\end{proof} | |
\begin{proposition} | |
$\ext^i(M,N)$ is a functor of $N$. | |
\end{proposition} | |
\begin{proof} | |
Obvious from the definition. | |
\end{proof} | |
Here is a harder statement. | |
\begin{proposition} | |
$\ext^i(M,N)$ is well-defined, independent of the projective | |
resolution $P_* | |
\to M$, and is in fact a contravariant additive functor of | |
$M$.\footnote{I.e. a map $M | |
\to M'$ induces $\ext^i(M', N) \to \ext^i(M,N)$.} | |
\end{proposition} | |
\begin{proof} | |
Omitted. We won't really need this, though; it requires more | |
theory about | |
chain complexes. | |
\end{proof} | |
\begin{proposition} | |
If $M$ is annihilated by some ideal $I \subset R$, then so is | |
$\ext^i(M,N)$ for | |
each $i$. | |
\end{proposition} | |
\begin{proof} | |
This is a consequence of the functoriality in $M$. If $x \in | |
I$,then $x: M \to | |
M$ is the zero map, so it induces the zero map on | |
$\ext^i(M,N)$.\end{proof} | |
\begin{proposition} | |
$\ext^i(M,N) = 0$ if $M$ projective and $i>0$. | |
\end{proposition} | |
\begin{proof} | |
In that case, one can use the projective resolution | |
\[ 0 \to M \to M \to 0. \] | |
Computing $\ext$ via this gives the result. | |
\end{proof} | |
\begin{proposition} | |
If there is an exact sequence | |
\[ 0 \to N' \to N \to N'' \to 0, \] | |
there is a long exact sequence of $\ext$ groups | |
\[ 0 \to \hom(M,N') \to \hom(M,N) \to \hom(M,N'') \to | |
\ext^1(M,N') \to | |
\ext^1(M,N) \to \dots \] | |
\end{proposition} | |
\begin{proof} | |
This proof will assume a little homological algebra. Choose a | |
projective | |
resolution $P_* \to M$. (The notation $P_*$ means the chain | |
complex $\dots \to | |
P_2 \to P_1 \to P_0$.) In general, homming out of $M$ is not | |
exact, but homming | |
out of a projective module is exact. For each $i$, we get an | |
exact sequence | |
\[ 0 \to \hom_R(P_i, N') \to \hom_R(P_i, N) \to \hom_R(P_i, | |
N'')\to 0, \] | |
which leads to an exact sequence of \emph{chain complexes} | |
\[ 0 \to \hom_R(P_*,N') \to \hom_R(P_*,N) \to \hom_R(P_*,N'') | |
\to 0 . \] | |
Taking the long exact sequence in homology gives the result. | |
\end{proof} | |
Much less obvious is: | |
\begin{proposition} | |
There is a long exact sequence in the $M$ variable. That is, a | |
short exact | |
sequence | |
\[ 0 \to M' \to M \to M'' \to 0 \] | |
leads a long exact sequence | |
\[ 0 \to \hom_R(M'', N) \to \hom_R(M,N) \to \hom_R(M', N) \to | |
\ext^1(M'', N) | |
\to \ext^1(M, N) \to \dots. \] | |
\end{proposition} | |
\begin{proof} | |
Omitted. | |
\end{proof} | |
We now can characterize projectivity: | |
\begin{corollary} | |
TFAE: | |
\begin{enumerate} | |
\item $M$ is projective. | |
\item $\ext^i(M,N) = 0$ for all $R$-modules $N$ and $i>0$. | |
\item $\ext^1(M,N)=0$ for all $N$. | |
\end{enumerate} | |
\end{corollary} | |
\begin{proof} | |
We have seen that 1 implies 2 because projective modules have | |
simple projective | |
resolutions. 2 obviously implies 3. Let's show that 3 implies | |
1.Choose a | |
projective module $P$ and a surjection $P \twoheadrightarrow M$ | |
with kernel | |
$K$. There is a short exact sequence $0 \to K \to P \to M \to | |
0$. The sequence | |
\[ 0 \to \hom(M,K) \to \hom(P,K) \to \hom(K,K) \to | |
\ext^1(M,K)=0\] | |
shows that there is a map $P \to K$ which restricts to the | |
identity $K \to K$. | |
The sequence $0 \to K \to P \to M \to 0$ thus splits, so $M$ is | |
a direct | |
summand in a projective module, so is projective. | |
\end{proof} | |
Finally, we note that there is another way of constructing | |
$\ext$. We | |
constructed them by choosing a projective resolution of $M$. But | |
you can also | |
do this by resolving $N$ by \emph{injective} modules. | |
\begin{definition} | |
An $R$-module $Q$ is \textbf{injective} if $\hom_R(-,Q)$ is an | |
exact (or, | |
equivalently, right-exact) functor. That is, if $M_0 \subset M$ | |
is an inclusion | |
of $R$-modules, then any map $M_0 \to Q$ can be extended to $M | |
\to Q$. | |
\end{definition} | |
If we are given $M,N$, and an injective resolution $N \to Q_*$, | |
we can look at | |
the chain complex $\left\{\hom(M,Q_i)\right\}$, i.e. the chain | |
complex | |
\[ 0 \to \hom(M, Q^0) \to \hom(M, Q^1) \to \dots \] | |
and we can consider the cohomologies. | |
\begin{definition} | |
We call these cohomologies | |
\[ \ext^i_R(M,N)' = \ker(\hom(M, Q^i) \to \hom(M, | |
Q^{i+1}))/\im(\hom(M, | |
Q^{i-1}) \to \hom(M, Q^i)). \] | |
\end{definition} | |
This is dual to the previous definitions, and it is easy to | |
check that the | |
properties that we couldn't verify for the previous $\ext$s are | |
true for the | |
$\ext'$'s. | |
Nonetheless: | |
\begin{theorem} | |
There are canonical isomorphisms: | |
\[ \ext^i(M,N)' \simeq \ext^i(M,N). \] | |
\end{theorem} | |
In particular, to compute $\ext$ groups, you are free either to | |
take a | |
projective resolution of $M$, or an injective resolution of | |
$N$.\begin{proof}[Idea of proof] | |
In general, it might be a good idea to construct a third more | |
complex | |
construction that resembles both. Given $M,N$ construct a | |
projective resolution | |
$P_* \to M$ and an injective resolution $N \to Q^*$. Having made | |
these choices, | |
we get a \emph{double complex} | |
\[ \hom_R(P_i, Q^j) \] | |
of a whole lot of $R$-modules. The claim is that in such a | |
situation, where | |
you have a double complex $C_{ij}$, you can | |
form an ordinary chain complex $C'$ | |
by adding along the diagonals. Namely, the $n$th term | |
is $C'_n = \bigoplus_{i+j=n} C_{ij}$. This \emph{total complex} | |
will receive a | |
map from the chain complex used to compute the $\ext$ groups | |
and a chain | |
complex used to compute the $\ext'$ groups. There are maps on | |
cohomology, | |
\[ \ext^i(M,N) \to H^i(C'_*), \quad \ext^i(M,N)' \to H^i(C'_*). | |
\] | |
The claim is that isomorphisms on | |
cohomology will be induced in each case. That will prove the | |
result, but we | |
shall not prove the claim. | |
\end{proof} | |
Last time we were talking about $\ext$ groups over commutative | |
rings. For $R$ a | |
commutative ring and $M,N$ $R$-modules, we defined an $R$-module | |
$\ext^i(M,N)$ for | |
each $i$, and proved various properties. We forgot to mention | |
one. | |
\begin{proposition} | |
If $R$ noetherian, and $M,N$ are finitely generated, | |
$\ext^i(M,N)$ is also finitely generated. | |
\end{proposition} | |
\begin{proof} | |
We can take a projective resolution $P_*$ of $M$ by finitely | |
generated free modules, $R$ being | |
noetherian. Consequently the complex $\hom(P_*, N)$ consists of | |
finitely | |
generated modules. Thus the cohomology is finitely generated, | |
and this cohomology | |
consists of the $\ext$ groups. | |
\end{proof} | |
\subsection{Application: Modules over DVRs} | |
\begin{definition} Let $M$ be a module over a domain $A$. We say that $M$ is \underline{torsion-free}, if for any nonzero $a \in A$, $a:M \to M$ is injective. We say that $M$ is \underline{torsion} if for any $m \in M$, there is nonzero $a \in A$ such that $am=0$. | |
\end{definition} | |
\begin{lemma} For any module finitely generated module $M$ over a Noetherian domain $A$, there is a short exact sequence | |
\[0 \to M_{tors} \to M \to M_{tors-free} \to 0\] | |
where $M_{tors}$ is killed by an element of $A$ and $M_{tors-free}$ is torsion-free. | |
\label{tors tors-free ses} | |
\end{lemma} | |
\begin{proof} This is because we may take $M_{tors}$ to be all the elements which are killed by a non-zero element of $A$. Then this is clearly a sub-module. Since $A$ is Noetherian, it is finitely generated, which means that it can be killed by one element of $A$ (take the product of the elements that kill the generators). Then it is easy to check that the quotient $M/M_{tors}$ is torsion-free. | |
\end{proof} | |
\begin{lemma} For $R$ a PID, a module $M$ over $R$ is flat if and only if it is torsion-free. | |
\label{PID means flat=tors free} | |
\end{lemma} | |
\begin{proof} This is the content of Problem 2 on the Midterm. | |
\end{proof} | |
Using this, we will classify modules over DVRs. | |
\begin{proposition} let $M$ be a finitely generated module over a DVR $R$. Then | |
\[M=M_{tors}\oplus R^{\oplus n},\] i.e, where $M_{tors}$ can be annihilated by $\pi^n$ for some $n$. | |
\end{proposition} | |
\begin{proof} | |
Set $M_{tors} \subset M$ be as in Lemma \ref{tors tors-free ses} so that $M/M_{tors}$ is torsion-free. Therefore, by Corollary \ref{DVR is PID} and Lemma \ref{PID means flat=tors free} we see that it is flat. But it is over a local ring, so that means that it is free. So we have $M/M_{tors}=R^{\oplus n}$ for some $n$. Furthermore, since $R^{\oplus n}$ is free, it is additionally projective, so the above sequence splits, so | |
\[M=M_{tors} \oplus R^{\oplus n}\] | |
as desired. | |
\end{proof} | |
There is nothing more to say about the free part, so let us discuss the torsion part in more detail. | |
\begin{lemma} Any finitely generated torsion module over a DVR is | |
\[\bigoplus R/\pi^nR.\] | |
\label{dvr fin gen tor module struct} | |
\end{lemma} | |
Before we prove this, let us give two examples: | |
\begin{enumerate} | |
\item Take $R=k[[t]]$, which is a DVR with maximal ideal (t). Thus, by the lemma, for a finitely generated torsion module $M$, $t:M \to M$ is a nilpotent operator. However, $k[[t]]/t^n$ is a Jordan block so we are exactly saying that linear transformations can be written in Jordan block form. | |
\item Let $R=\mathbb{Z}_p$. Here the lemma implies that finitely generated torsion modules over $\mathbb{Z}_p$ can be written as a direct sum of $p$-groups. | |
\end{enumerate} | |
Now let us proceed with the proof of the lemma. | |
\begin{proof}[Proof of Lemma \ref{dvr fin gen tor module struct}] Let $n$ be the minimal integer such that $\pi^n$ kills $M$. This means that $M$ is a module over $R_n=R/\pi^nR$, and also there is an element $m \in M$, and an injective map $R_n \hookrightarrow M$, because we may choose $m$ to be an element which is not annihilated by $\pi^{n-1}$, and then take the map to be $1 \mapsto m$. | |
Proceeding by induction, it suffices to show that the above map $R_n \hookrightarrow M$ splits. But for this it suffices that $R_n$ is an injective module over itself. This property of rings is called the Frobenius property, and it is very rare. We will write this as a lemma. | |
\begin{lemma} $R_n$ is injective as a module over itself. | |
\label{Rn Frobenius} | |
\end{lemma} | |
\begin{proof}[Proof of Lemma \ref{Rn Frobenius}] Note that a module $M$ over a ring $R$ is injective if and only if for any ideal $I \subset R$, $\Ext^1(R/I,M)=0$. This was shown on Problem Set 8, Problem 2a. | |
Thus we wish to show that for any ideal $I$, $\Ext^1_{R_n}(R_n/I,R_n)=0$. Note that since $R$ is a DVR, we know that it is a PID, and also any element has the form $r=\pi^kr_0$ for some $k \geq 0$ and some $r_0$ invertible. Then all ideals in $R$ are of the form $(\pi^k)$ for some $k$, so all ideals in $R_n$ are also of this form. Therefore, $R_n/I=R_m$ for some $m \leq n$, so it suffices to show that for $m \leq n$, $\Ext^1_{R_n}(R_m,R_n)=0$. | |
But note that we have short exact sequence | |
\[ 0 \to R_{n-m} \to^{\pi^m \cdot} R_n \to R_m \to 0\] | |
which gives a corresponding long exact sequence of $\Ext$s | |
\[0 \to \Hom_{R_n}(R_m,R_n) \to \Hom_{R_n}(R_n,R_n) \to^\hearts \Hom_{R_n}(R_{n-m},R_n)\] | |
\[\to \Ext^1_{R_n}(R_m,R_n) \to \Ext^1_{R_n}(R_n,R_n) \to \cdots\] | |
But note that any map of $R_n$ modules, $R_{n-m} \to R_n$, must map $1 \in R_{n-m}$ to an element which is killed by $\pi^{n-m}$, which means it must be a multiple of $\pi^m$, so say is is $\pi^ma$. Then the map is | |
\[r \mapsto \pi^mar,\] | |
which is the image of the map | |
\[[r \mapsto ar] \in \Hom_{R_n}(R_n,R_n).\] | |
Thus, $\hearts$ is surjective. | |
Also note that $R_n$ is projective over itself, so $\Ext^1_{R_n}(R_n,R_n)=0$. This, along with the surjectivity of $\hearts$ shows that | |
\[\Ext^1_{R_n}(R_m,R_n)=0\] | |
as desired. | |
\end{proof} | |
As mentioned earlier, this lemma concludes our proof of Lemma \ref{dvr fin gen tor module struct} as well. | |
\end{proof} | |