Datasets:
Tasks:
Text Generation
Modalities:
Text
Sub-tasks:
language-modeling
Languages:
English
Size:
100K - 1M
License:
\chapter{Three important functors} | |
There are three functors that will be integral to our study of commutative | |
algebra in the future: localization, the tensor product, and $\hom$. | |
While localization is an \emph{exact} functor, the tensor product and $\hom$ | |
are not. The failure of exactness in those cases leads to the theory of | |
flatness and projectivity (and injectivity), and eventually the \emph{derived functors} | |
$\mathrm{Tor}$ and $\mathrm{Ext}$ that crop up in commutative algebra. | |
\section{Localization} | |
Localization is the process of making invertible a collection of elements in a | |
ring. It is a generalization of the process of forming a quotient field of an | |
integral domain. | |
\subsection{Geometric intuition} | |
We first start off with some of the geometric intuition behind the idea of | |
localization. Suppose we have a Riemann surface $X$ (for example, the Riemann | |
sphere). Let $A(U)$ be the ring of holomorphic functions over some neighborhood | |
$U\subset X$. Now, for holomorphicity to hold, all that is required is | |
that a function doesn't have a pole inside of $U$, thus when $U=X$, this | |
condition is the strictest and as $U$ gets smaller functions begin to show | |
up that may not arise from the restriction of a holomorphic function over | |
a larger domain. For example, if we want to study holomorphicity ``near a | |
point $z_0$'' all that we should require is that the function doesn't pole at | |
$z_0$. This means that we should consider quotients of holomorphic functions | |
$f/g$ where $g(z_0)\neq 0$. This process of inverting a collection of elements | |
is expressed through the algebraic construction known as ``localization.'' | |
\subsection{Localization at a multiplicative subset} | |
Let $R$ be a commutative ring. | |
We start by constructing the notion of \emph{localization} in the most general | |
sense. | |
We have already implicitly used this definition, but nonetheless, we make it | |
formally: | |
\begin{definition} \label{multset} | |
A subset $S \subset R$ is a \textbf{multiplicative subset} if $1 \in S$ and | |
if $x,y \in S$ implies $xy \in S$. | |
\end{definition} | |
We now define the notion of \emph{localization}. Formally, this means | |
inverting things. | |
This will give us a functor from $R$-modules to $R$-modules. | |
\begin{definition} | |
If $M$ is an $R$-module, we define the module $S^{-1}M$ as the set of formal | |
fractions | |
\[ \left\{m/s, m \in M, s \in S\right\} \] | |
modulo an equivalence relation: where $m/s \sim m'/s'$ if and only if | |
\[ t( s'm - m's ) = 0 \] | |
for some $t \in S$. The reason we need to include the $t$ in the definition | |
is that otherwise the | |
relation would not be transitive (i.e. would not be an | |
equivalence relation). | |
\end{definition} | |
So two fractions agree if they agree when clearing denominators and | |
multiplication. | |
It is easy to check that this is indeed an equivalence relation. Moreover | |
$S^{-1}M$ is an abelian group with the usual addition of fractions | |
\[ \frac{m}{s}+\frac{m'}{s'} = \frac{s'm + sm'}{ss'} \] | |
and it is easy to check that this is a legitimate abelian group. | |
\begin{definition} | |
Let $M$ be an $R$-module and $S \subset R$ a multiplicative subset. | |
The abelian group $S^{-1}M$ is naturally an $R$-module. We define | |
\[ x(m/s) = (xm)/s, \quad x \in R. \] | |
It is easy to check that this is well-defined and makes it into a module. | |
Finally, we note that localization is a \emph{functor} from the category of | |
$R$-modules to itself. Indeed, given $f: M \to N$, there is a naturally | |
induced map $S^{-1}M \stackrel{S^{-1}f}{\to} S^{-1}N$. | |
\end{definition} | |
We now consider the special case when the localized module is the initial ring | |
itself. | |
Let $M = R$. Then $S^{-1}R$ is an $R$-module, and it is in fact a commutative | |
ring in its own right. The ring structure is quite tautological: | |
\[ (x/s)(y/s') = (xy/ss'). \] | |
There is a map $R \to S^{-1}R$ sending $x \to x/1$, which is a | |
ring-homomorphism. | |
\begin{definition} | |
For $S \subset R$ a multiplicative set, the localization $S^{-1}R$ is a | |
commutative ring as above. In fact, it is an $R$-algebra; there is a natural | |
map $\phi: R \to S^{-1}R$ sending $r \to r/1$. | |
\end{definition} | |
We can, in fact, describe $\phi: R \to S^{-1}R$ by a \emph{universal | |
property}. Note | |
that for each $s \in S$, $\phi(s)$ is invertible. This is because $\phi(s) = | |
s/1$ which has a multiplicative inverse $1/s$. This property characterizes | |
$S^{-1}R$. | |
For any commutative ring $B$, $\hom(S^{-1}R, B)$ is naturally isomorphic to the | |
subset of $\hom(R,B)$ that send $S$ to units. The map takes $S^{-1}R \to B$ to | |
the pull-back $R \to S^{-1}R \to B$. The proof of this is very simple. | |
Suppose that $f: R \to B$ is such that $f(s) \in B$ is invertible for each $s | |
\in S$. Then we must define $S^{-1}R \to B$ by sending $r/s$ to | |
$f(r)f(s)^{-1}$. It is easy to check that this is well-defined and that the | |
natural isomorphism as claimed is true. | |
Let $R$ be a ring, $M$ an $R$-module, $S \subset R$ a multiplicatively closed | |
subset. We defined a ring of fractions $S^{-1}R$ and an $R$-module $S^{-1}M$. | |
But in fact this is a module over the ring $S^{-1}R$. | |
We just multiply $(x/t)(m/s) = (xm/st)$. | |
In particular, localization at $S$ gives a \emph{functor} from $R$-modules to | |
$S^{-1}R$-modules. | |
\begin{exercise} | |
Let $R$ be a ring, $S$ a multiplicative subset. Let $T$ be the $R$-algebra | |
$R[\left\{x_s\right\}_{s \in S}]/( \left\{sx_s - 1\right\})$. This is the | |
polynomial ring in the variables $x_s$, one for each $s \in S$, modulo the | |
ideal generated by $sx_s = 1$. Prove that this $R$-algebra is naturally | |
isomorphic to $S^{-1}R$, using the universal property. | |
\end{exercise} | |
\begin{exercise} Define a functor $\mathbf{Rings} \to \mathbf{Sets}$ sending | |
a ring to | |
its set of units, and show that it is corepresentable (use $\mathbb{Z}[X, | |
X^{-1}]$). | |
\end{exercise} | |
\subsection{Local rings} | |
A special case of great importance in the future is when the multiplicative | |
subset is the complement of a prime ideal, and we study this in the present | |
subsection. Such localizations will be ``local rings'' and geometrically | |
correspond to the process of zooming at a point. | |
\begin{example} | |
Let $R$ be an integral domain and let $S = R - \left\{0\right\}$. This is a | |
multiplicative subset because $R$ is a domain. In this case, $S^{-1}R$ is just | |
the ring of fractions by allowing arbitrary nonzero denominators; it is a | |
field, and is called the \textbf{quotient field}. The most familiar example is | |
the construction of $\mathbb{Q}$ as the quotient field of $\mathbb{Z}$. | |
\end{example} | |
We'd like to generalize this example. | |
\begin{example} | |
Let $R$ be arbitrary and $\mathfrak{p}$ is a prime ideal. This means that $1 | |
\notin \mathfrak{p}$ and $x,y \in R - \mathfrak{p}$ implies that $xy \in R - | |
\mathfrak{p}$. Hence, the complement $S = R- \mathfrak{p}$ is multiplicatively | |
closed. We get a ring $S^{-1}R$. | |
\begin{definition} | |
This ring is denoted $R_{\mathfrak{p}}$ and is called the \textbf{localization | |
at $\mathfrak{p}$.} If $M$ is an $R$-module, we write $M_{\mathfrak{p}}$ for | |
the localization of $M$ at $R - \mathfrak{p}$. | |
\end{definition} | |
This generalizes the previous example (where $\mathfrak{p} = (0)$). | |
\end{example} | |
There is a nice property of the rings $R_{\mathfrak{p}}$. To elucidate this, | |
we start with a lemma. | |
\begin{lemma} | |
Let $R$ be a nonzero commutative ring. The following are equivalent: | |
\begin{enumerate} | |
\item $R$ has a unique maximal ideal. | |
\item If $x \in R$, then either $x$ or $1-x$ is invertible. | |
\end{enumerate} | |
\end{lemma} | |
\begin{definition} | |
In this case, we call $R$ \textbf{local}. A local ring is one with a unique | |
maximal ideal. | |
\end{definition} | |
\begin{proof}[Proof of the lemma] | |
First we prove $(2) \implies (1)$. | |
Assume $R$ is such that for | |
each $x$, either $x$ or $1-x$ is invertible. We will find the maximal ideal. | |
Let $\mathfrak{M} $ be the collection of noninvertible elements of $R$. This is | |
a subset of $R$, not containing $1$, and it is closed under multiplication. | |
Any proper ideal must be a subset of $\mathfrak{M}$, because otherwise that | |
proper ideal would contain an invertible element. | |
We just need to check that $\mathfrak{M}$ is closed under addition. | |
Suppose to the | |
contrary that $x, y \in \mathfrak{M}$ but $x+y$ is invertible. We get (with | |
$a = x/(x+y)$) | |
\[ 1 = \frac{x}{x+y} + \frac{y}{x+y} =a+(1-a). \] | |
Then one of $a,1-a$ is invertible. So either $x(x+y)^{-1}$ or $y(x+y)^{-1}$ is | |
invertible, which implies that either $x,y$ is invertible, contradiction. | |
Now prove the reverse direction. Assume $R$ has a unique maximal ideal | |
$\mathfrak{M}$. We claim that $\mathfrak{M}$ consists precisely of the | |
noninvertible elements. To see this, first note that $\mathfrak{M}$ | |
can't contain any invertible elements since it is proper. Conversely, suppose | |
$x$ is not invertible, i.e. $(x) \subsetneq R$. Then $(x)$ is contained in a | |
maximal ideal by \rref{anycontainedinmaximal}, so $(x) \subset | |
\mathfrak{M}$ since $\mathfrak{M}$ is unique among maximal ideals. | |
Thus $x \in \mathfrak{M}$. | |
Suppose $x \in R$; we can write $1 = x + (1-x)$. Since $1 \notin \mathfrak{M}$, | |
one of $x, 1-x$ must not be in $\mathfrak{M}$, so one of those must not be | |
invertible. So $(1) \implies (2)$. The lemma is proved. | |
\end{proof} | |
Let us give some examples of local rings. | |
\begin{example} | |
Any field is a local ring because the unique maximal ideal is $(0)$. | |
\end{example} | |
\begin{example} | |
Let $R$ be any commutative ring and $\mathfrak{p}\subset R$ a prime ideal. Then | |
$R_{\mathfrak{p}}$ is a local ring. | |
We state this as a result. | |
\begin{proposition} | |
$R_{\mathfrak{p}}$ is a local ring if $\mathfrak{p}$ is prime.\end{proposition} | |
\begin{proof} | |
Let $\mathfrak{m} \subset R_{\mathfrak{p}}$ consist of elements $x/s$ for $x | |
\in \mathfrak{p}$ and $s \in R - \mathfrak{p}$. It is left as an exercise | |
(using the primality of $\mathfrak{p}$) to | |
the reader to see that whether the numerator belongs to $\mathfrak{p}$ is | |
\emph{independent} of the representation $x/s$ used for it. | |
Then I claim that $\mathfrak{m}$ is the | |
unique maximal ideal. First, note that $\mathfrak{m}$ is | |
an ideal; this is evident since the numerators form an ideal. If $x/s, y/s'$ | |
belong to $\mathfrak{m}$ with appropriate expressions, then | |
the numerator of | |
\[ \frac{xs'+ys}{ss'} \] | |
belongs to $\mathfrak{p}$, so this sum belongs to $\mathfrak{m}$. Moreover, | |
$\mathfrak{m}$ is a proper ideal because $\frac{1}{1}$ is not of the | |
appropriate form. | |
I claim that $\mathfrak{m}$ contains all other proper ideals, which will imply | |
that it is the unique maximal ideal. Let $I \subset R_{\mathfrak{p}}$ be any | |
proper ideal. Suppose $x/s \in I$. We want to prove $x/s \in \mathfrak{m}$. | |
In other words, we have to show $x \in \mathfrak{p}$. But if not $x/s$ would be | |
invertible, and $I = (1)$, contradiction. This proves locality. | |
\end{proof} | |
\end{example} | |
\begin{exercise} | |
Any local ring is of the form $R_{\mathfrak{p}}$ for some ring $R$ and for | |
some prime ideal $\mathfrak{p} \subset R$. | |
\end{exercise} | |
\begin{example} | |
Let $R = \mathbb{Z}$. This is not a local ring; the maximal ideals are given by | |
$(p)$ for $p$ prime. We can thus construct the localizations | |
$\mathbb{Z}_{(p)}$ of all fractions $a/b \in \mathbb{Q}$ where $b \notin (p)$. | |
Here $\mathbb{Z}_{(p)}$ consists of all rational numbers that don't have | |
powers of $p$ in the denominator. | |
\end{example} | |
\begin{exercise} | |
A local ring has no idempotents other than $0$ and $1$. (Recall that $e \in R$ | |
is \emph{idempotent} if $e^2 = e$.) In particular, the product of two rings is | |
never local. | |
\end{exercise} | |
It may not yet be clear why localization is such a useful process. It turns | |
out that many problems can be checked on the localizations at prime (or even | |
maximal) ideals, so certain proofs can reduce to the case of a local ring. | |
Let us give a small taste. | |
\begin{proposition} | |
Let $f: M \to N$ be a homomorphism of $R$-modules. Then $f$ is injective if | |
and only if for every maximal ideal $\mathfrak{m} \subset R$, we have that | |
$f_{\mathfrak{m}}: M_{\mathfrak{m}} \to N_{\mathfrak{m}}$ is injective. | |
\end{proposition} | |
Recall that, by definition, $M_{\mathfrak{m}}$ is the localization at $R - | |
\mathfrak{m}$. | |
There are many variants on this (e.g. replace with surjectivity, bijectivity). | |
This is a general observation that lets you reduce lots of commutative algebra | |
to local rings, which are easier to work with. | |
\begin{proof} | |
Suppose first that each $f_{\mathfrak{m}}$ is injective. I claim that $f$ is | |
injective. Suppose $x \in M - \left\{0\right\}$. We must show that $f(x) \neq | |
0$. If $f(x)=0$, then $f_{\mathfrak{m}}(x)=0$ for every maximal ideal | |
$\mathfrak{m}$. Then by | |
injectivity it follows that $x$ maps to zero in each $M_{\mathfrak{m}}$. | |
We would now like to get a contradiction. | |
Let $I = \left\{ a \in R: ax = 0 \in M \right\}$. This is proper since $x \neq | |
0$. So $I$ is contained in some maximal ideal $\mathfrak{m}$. Then $x$ | |
maps to zero in $M_{\mathfrak{m}}$ by the previous paragraph; this means that | |
there is $s \in R - \mathfrak{m}$ with $sx = 0 \in M$. But $s \notin I$, | |
contradiction. | |
Now let us do the other direction. Suppose $f$ is injective and $\mathfrak{m}$ | |
a maximal ideal; we prove $f_{\mathfrak{m}}$ injective. Suppose | |
$f_{\mathfrak{m}}(x/s)=0 \in N_{\mathfrak{m}}$. This means that $f(x)/s=0$ in | |
the localized module, so that $f(x) \in M$ is killed by some $t \in R - | |
\mathfrak{m}$. We thus have $f(tx) = t(f(x)) = 0 \in M$. This means that $tx | |
= 0 \in M$ since $f$ is injective. But this in turn means that $x/s = 0 \in | |
M_{\mathfrak{m}}$. This is what we wanted to show. | |
\end{proof} | |
\subsection{Localization is exact} | |
Localization is to be thought of as a very mild procedure. | |
The next result says how inoffensive localization is. This result is a key | |
tool in reducing problems to the local case. | |
\begin{proposition} | |
Suppose $f: M \to N, g: N \to P$ and $M \to N \to P$ is exact. Let $S \subset | |
R$ be multiplicatively closed. Then | |
\[ S^{-1}M \to S^{-1}N \to S^{-1}P \] | |
is exact. | |
\end{proposition} | |
Or, as one can alternatively express it, localization is an \emph{exact | |
functor.} | |
Before proving it, we note a few corollaries: | |
\begin{corollary} | |
If $f: M \to N$ is surjective, then $S^{-1}M \to S^{-1}N$ is too. | |
\end{corollary} | |
\begin{proof} | |
To say that $A \to B$ is surjective is the same as saying that $A \to B \to 0$ | |
is exact. From this the corollary is evident. | |
\end{proof} | |
Similarly: | |
\begin{corollary} | |
If $f: M \to N$ is injective, then $S^{-1}M \to S^{-1}N$ is too. | |
\end{corollary} | |
\begin{proof} | |
To say that $A \to B$ is injective is the same as saying that $0 \to A \to B $ | |
is exact. From this the corollary is evident. | |
\end{proof} | |
\begin{proof}[Proof of the proposition] We adopt the notation of the | |
proposition. | |
If the composite $g\circ f$ is zero, clearly the localization $S^{-1}M \to | |
S^{-1}N \to S^{-1}P$ is zero too. | |
Call the maps $S^{-1}M \to S^{-1}N, S^{-1}N \to S^{-1}P$ as $\phi, \psi$. We | |
know that $\psi \circ \phi = 0$ so $\ker(\psi) \supset \im(\phi)$. Conversely, | |
suppose something belongs to $\ker(\psi). $ This can be written as a fraction | |
\[ x/s \in \ker(\psi) \] | |
where $x \in N, s \in S$. This is mapped to | |
\[ g(x)/s \in S^{-1}P, \] | |
which we're assuming is zero. This means that there is $t \in S$ with $tg(x) = | |
0 \in P$. This means that $g(tx)=0$ as an element of $P$. But $tx \in N$ and | |
its image of $g$ vanishes, so $tx$ must come from something in $M$. In | |
particular, | |
\[ tx = f(y) \ \text{for some} \ y \in M. \] | |
In particular, | |
\[ \frac{x}{s} = \frac{tx}{ts} = \frac{f(y)}{ts} = \phi( y/ts) \in \im(\phi). | |
\] | |
This proves that anything belonging to the kernel of $\psi$ lies in | |
$\im(\phi)$. | |
\end{proof} | |
\subsection{Nakayama's lemma} | |
We now state a very useful criterion for determining when a module over a | |
\emph{local} ring is zero. | |
\begin{lemma}[Nakayama's lemma] \label{nakayama} If $R$ is a local ring with | |
maximal ideal | |
$\mathfrak{m}$. Let $M$ be a finitely generated $R$-module. If | |
$\mathfrak{m}M = M$, then $M = 0$. | |
\end{lemma} | |
Note that $\mathfrak{m}M$ is the submodule generated by products of | |
elements of $\mathfrak{m}$ and $M$. | |
\begin{remark} | |
Once one has the theory of the tensor product, this equivalently states that | |
if $M$ is finitely generated, then | |
\[ M \otimes_R R/\mathfrak{m} = M/\mathfrak{m}M \neq 0. \] | |
So to prove that a finitely generated module over a local ring is zero, you | |
can reduce to studying the reduction to $R/\mathfrak{m}$. This is thus a very | |
useful criterion. | |
\end{remark} | |
Nakayama's lemma highlights why it is so useful to work over a local ring. | |
Thus, it is useful to reduce questions about general rings to questions about | |
local rings. | |
Before proving it, we note a corollary. | |
\begin{corollary} | |
Let $R$ be a local ring with maximal ideal $\mathfrak{m}$, and $M$ a finitely | |
generated module. If $N \subset M$ is a submodule such that $N + | |
\mathfrak{m}N = | |
M$, then $N=M$. | |
\end{corollary} | |
\begin{proof} | |
Apply Nakayama above (\cref{nakayama}) to $M/N$. | |
\end{proof} | |
We shall prove more generally: | |
\begin{proposition} | |
Suppose $M$ is a finitely generated $R$-module, $J \subset R$ an ideal. | |
Suppose $JM = M$. Then there is $a \in 1+J$ such that $aM = 0$. | |
\end{proposition} | |
If $J$ is the maximal ideal of a local ring, then $a$ is a unit, so that $M=0$. | |
\begin{proof} | |
Suppose $M$ is generated by $\left\{x_1, \dots, x_n\right\} \subset M$. This | |
means that every element of $M$ is a linear combination of elements of | |
$x_i$. However, each $x_i \in JM$ by assumption. In particular, each | |
$x_i$ can be written as | |
\[ x_i = \sum a_{ij} x_j, \ \mathrm{where} \ a_{ij} \in \mathfrak{m}. \] | |
If we let $A$ be the matrix $\left\{a_{ij}\right\}$, then $A$ sends the | |
vector $(x_i)$ into itself. In particular, $I-A$ kills | |
the vector $(x_i)$. | |
Now $I-A$ is an $n$-by-$n$ matrix in the ring $R$. We could, of course, | |
reduce everything modulo $J$ to get the identity; this is | |
because $A$ consists of elements of $J$. It follows that the | |
determinant must be congruent to $1$ modulo $J$. | |
In particular, $a=\det (I - A)$ lies in $1+J$. | |
Now by familiar linear algebra, $aI$ can be represented as the product of $A$ | |
and the matrix of cofactors; in particular, $aI$ annihilates the vector | |
$(x_i)$, so that $aM=0$. | |
\end{proof} | |
Before returning to the special case of local rings, we observe the following | |
useful fact from ideal theory: | |
\begin{proposition} \label{idempotentideal} | |
Let $R$ be a commutative ring, $I \subset R$ a finitely generated ideal such that $I^2 = I$. | |
Then $I$ is generated by an idempotent element. | |
\end{proposition} | |
\begin{proof} | |
We know that there is $x \in 1+I$ such that $xI =0$. If $x = 1+y, y \in I$, it | |
follows that | |
\[ yt = t \] | |
for all $t \in I$. In particular, $y$ is idempotent and $(y) = I$. | |
\end{proof} | |
\begin{exercise} | |
\rref{idempotentideal} fails if the ideal is not finitely generated. | |
\end{exercise} | |
\begin{exercise} | |
Let $M$ be a finitely generated module over a ring $R$. Suppose $f: M \to M$ | |
is a surjection. Then $f$ is an isomorphism. To see this, consider $M$ as a | |
module over $R[t]$ with $t$ acting by $f$; since $(t)M = M$, argue that there | |
is a polynomial $Q(t) \in R[t]$ such that $Q(t)t$ acts as the identity on | |
$M$, i.e. $Q(f)f=1_M$. | |
\end{exercise} | |
\begin{exercise} | |
Give a counterexample to the conclusion of Nakayama's lemma when the module is | |
not finitely generated. | |
\end{exercise} | |
\begin{exercise} | |
Let $M$ be a finitely generated module over the ring $R$. Let $\mathfrak{I}$ | |
be the Jacobson | |
radical of $R$ (cf. \rref{Jacobson}). If $\mathfrak{I} M = M$, | |
then $M = | |
0$. | |
\end{exercise} | |
\begin{exercise}[A converse to Nakayama's lemma] | |
Suppose conversely that $R$ is a ring, and $\mathfrak{a} \subset R$ an ideal | |
such that $\mathfrak{a} M \neq M$ for every nonzero finitely generated | |
$R$-module. Then $\mathfrak{a}$ is contained in every maximal ideal of $R$. | |
\end{exercise} | |
\begin{exercise} | |
Here is an alternative proof of Nakayama's lemma. Let $R$ be local with | |
maximal ideal $\mathfrak{m}$, and let $M$ be a finitely generated module with | |
$\mathfrak{m}M = M$. Let $n$ be the minimal number of generators for $M$. If | |
$n>0$, pick generators $x_1, \dots, x_n$. Then write $x_1 = a_1 x_1 + \dots + | |
a_n x_n$ where each $a_i \in \mathfrak{m}$. Deduce that $x_1$ is in the | |
submodule generated by the $x_i, i \geq 2$, so that $n$ was not actually | |
minimal, contradiction. | |
\end{exercise} | |
Let $M, M'$ be finitely generated modules over a local ring $(R, | |
\mathfrak{m})$, and let $\phi: M \to M'$ be a homomorphism of modules. Then | |
Nakayama's lemma gives a criterion for $\phi$ to be a surjection: namely, the | |
map $\overline{\phi}: M/\mathfrak{m}M \to M'/\mathfrak{m}M'$ must be a surjection. | |
For injections, this is false. For instance, if $\phi$ is multiplication by any element of | |
$\mathfrak{m}$, then $\overline{\phi}$ is zero but $\phi$ may yet be injective. | |
Nonetheless, we give a criterion for a map of \emph{free} modules over a local ring to | |
be a \emph{split} injection. | |
\begin{proposition} \label{splitcriterion1} | |
Let $R$ be a local ring with maximal ideal $\mathfrak{m}$. Let $F, F'$ be two | |
finitely generated free $R$-modules, and let $\phi: F \to F'$ be a homomorphism. | |
Then $\phi$ is a split injection if and only if the reduction $\overline{\phi}$ | |
\[ F/\mathfrak{m}F \stackrel{\overline{\phi}}{\to} F'/\mathfrak{m}F' \] | |
is an injection. | |
\end{proposition} | |
\begin{proof} | |
One direction is easy. If $\phi$ is a split injection, then it has a left | |
inverse | |
$\psi: F' \to F$ such that $\psi \circ \phi = 1_F$. The reduction of $\psi$ as a | |
map $F'/\mathfrak{m}F' \to F/\mathfrak{m}F$ is a left inverse to | |
$\overline{\phi}$, which is thus injective. | |
Conversely, suppose $\overline{\phi}$ injective. Let $e_1, \dots, e_r$ be a | |
``basis'' for $F$, and let $f_1, \dots, f_r$ be the images under $\phi$ in | |
$F'$. Then the reductions $\overline{f_1}, \dots, \overline{f_r}$ are linearly | |
independent in the $R/\mathfrak{m}$-vector space $F'/\mathfrak{m}F'$. Let us | |
complete this to a basis of $F'/\mathfrak{m}F'$ by adding elements | |
$\overline{g_1}, \dots, \overline{g_s} \in F'/\mathfrak{m}F'$, which we can | |
lift to elements $g_1, \dots, g_s \in F'$. It is clear that $F'$ has rank $r+s $ | |
since its reduction $F'/\mathfrak{m}F'$ does. | |
We claim that the set $\left\{f_1, \dots, f_r, g_1, \dots, g_s\right\}$ is a | |
basis for $F'$. Indeed, we have a map | |
\[ R^{r+s} \to F' \] | |
of free modules of rank $r+s$. It can be expressed as an $r+s$-by-$r+s$ matrix | |
$M$; we need to show that $M$ is invertible. But if we reduce modulo | |
$\mathfrak{m}$, it is invertible since the reductions of $f_1, \dots, f_r, | |
g_1, \dots, g_s$ form a basis of $F'/\mathfrak{m}F'$. | |
Thus the determinant of $M$ is not in $\mathfrak{m}$, so by locality it is | |
invertible. | |
The claim about $F'$ is thus proved. | |
We can now define the left inverse $F' \to F$ of $\phi$. Indeed, given $x \in F'$, | |
we can write it uniquely as a linear combination $\sum a_i f_i + \sum b_j g_j$ | |
by the above. We define $\psi(\sum a_i f_i + \sum b_j g_j) = \sum a_i e_i \in | |
F$. It is clear that this is a left inverse | |
\end{proof} | |
We next note a slight strenghtening of the above result, which is sometimes | |
useful. Namely, the first module does not have to be free. | |
\begin{proposition} | |
Let $R$ be a local ring with maximal ideal $\mathfrak{m}$. Let $M, F$ be two | |
finitely generated $R$-modules with $F$ free, and let $\phi: M \to F'$ be a homomorphism. | |
Then $\phi$ is a split injection if and only if the reduction $\overline{\phi}$ | |
\[ M/\mathfrak{m}M \stackrel{\overline{\phi}}{\to} F/\mathfrak{m}F \] | |
is an injection. | |
\end{proposition} | |
It will in fact follow that $M$ is itself free, because $M$ is projective (see | |
\cref{} below) as it is a direct summand of a free module. | |
\begin{proof} | |
Let $L$ be a ``free approximation'' to $M$. | |
That is, choose a basis $\overline{x_1}, \dots, \overline{x_n}$ for $M/\mathfrak{m}M$ (as an $R/\mathfrak{m}$-vector | |
space) and lift this to elements $x_1, \dots, x_n \in M$. Define a map | |
\[ L = R^n \to M \] | |
by sending the $i$th basis vector to $x_i$. | |
Then $L/\mathfrak{m} L \to M/\mathfrak{m}M$ is an isomorphism. | |
By Nakayama's lemma, | |
$L \to M$ is surjective. | |
Then the composite map | |
$L \to M \to F$ is such that the $L/\mathfrak{m}L \to F/\mathfrak{m}F$ is injective, so | |
$L \to F$ is a split injection (by \cref{splitcriterion1}). | |
It follows that we can find a splitting $F \to L$, which when composed with $L | |
\to M$ is a splitting of $M \to F$. | |
\end{proof} | |
\begin{exercise} | |
Let $A$ be a local ring, and $B$ a ring which is finitely generated and free as an | |
$A$-module. Suppose $A \to B$ is an injection. Then $A \to B$ is a \emph{split | |
injection.} (Note that any nonzero morphism mapping out of a field is | |
injective.) | |
\end{exercise} | |
\section{The functor $\hom$} | |
In any category, the morphisms between two objects form a | |
set.\footnote{Strictly speaking, this may depend on your set-theoretic | |
foundations.} In many | |
categories, however, the hom-sets have additional structure. For instance, | |
the hom-sets | |
between abelian groups are themselves abelian groups. The same situation holds | |
for the category of modules over a commutative ring. | |
\begin{definition} | |
Let $R$ be a commutative ring and $M,N$ to be $R$-modules. We write | |
$\hom_R(M,N)$ for | |
the set of all $R$-module homomorphisms $M \to N$. | |
$\hom_R(M,N)$ is an $R$-module because one can add homomorphisms $f,g: M | |
\to N$ by adding | |
them pointwise: if $f,g$ are homomorphisms $M \to N$, define $f+g: M \to N$ via | |
\( (f+g)(m) = f(m)+g(m); \) | |
similarly, one can multiply homomorphisms $f: M \to N$ by elements $ a \in | |
R$: one sets | |
\( (af)(m) = a(f(m)). \) | |
\end{definition} | |
Recall that in any category, the hom-sets are \emph{functorial}. For instance, | |
given $f: N \to N'$, post-composition with $f$ defines a map $\hom_R(M,N) \to | |
\hom_R(M,N')$ for any $M$. | |
Similarly precomposition gives a natural map $\hom_R(N', M) \to \hom_R(N, M)$. | |
In particular, we get a bifunctor $\hom$, contravariant in the first variable | |
and covariant in the second, of $R$-modules into $R$-modules. | |
\subsection{Left-exactness of $\hom$} | |
We now discuss the exactness properties of this construction of forming | |
$\hom$-sets. The following result is basic and is, in fact, a reflection of | |
the universal property of the kernel. | |
\begin{proposition} \label{homcovleftexact} | |
If $M$ is an $R$-module, then the functor | |
\[ N \to \hom_R(M,N) \] | |
is left exact (but \emph{not exact} in general). | |
\end{proposition} | |
This means that if | |
\[ 0 \to N' \to N \to N'' \] | |
is exact, | |
then | |
\[ 0 \to \hom_R(M, N') \to \hom_R(M, N) \to \hom_R(M, N'') \] | |
is exact as well. | |
\begin{proof} | |
First, we have to show that the map | |
$\hom_R(M,N') \to \hom_R(M,N)$ is injective; this is because $N' \to N$ is | |
injective, and composition with $N' \to N$ can't kill any nonzero $M \to N'$. | |
Similarly, exactness in the middle can be checked easily, and follows from | |
\rref{univpropertykernel}; it states simply that a map $M \to N$ has | |
image landing inside $N'$ (i.e. factors through $N'$) if and only if it | |
composes to zero in $N''$. | |
\end{proof} | |
\newcommand{\ol}[1]{\mathbf{#1}} | |
This functor $\hom_R(M, \cdot)$ is not exact in general. Indeed: | |
\begin{example} | |
Suppose $R = \mathbb{Z}$, and consider the $R$-module (i.e. abelian group) | |
$M = \mathbb{Z}/2\mathbb{Z}$. There is a short exact | |
sequence | |
\[ 0 \to 2\mathbb{Z} \to \mathbb{Z} \to \mathbb{Z}/2\mathbb{Z} \to 0. \] | |
Let us apply $\hom_R(M, \cdot)$. We get a \emph{complex} | |
\[ 0 \to \hom(\mathbb{Z}/2\mathbb{Z}, 2\mathbb{Z}) \to | |
\hom(\mathbb{Z}/2\mathbb{Z}, \mathbb{Z}) \to \hom(\mathbb{Z}/2\mathbb{Z}, | |
\mathbb{Z}/2\mathbb{Z}) \to 0. \] | |
The second-to-last term is $\mathbb{Z}/2\mathbb{Z}$; everything else is | |
zero. Thus the sequence is not exact, and in particular the functor | |
$\hom_{\mathbb{Z}}(\mathbb{Z}/2, -)$ is not an exact functor. | |
\end{example} | |
We have seen that homming out of a module is left-exact. Now, we see the same | |
for homming \emph{into} a module. | |
\begin{proposition} \label{homcontleftexact} | |
If $M$ is a module, then $\hom_R(-,M)$ is a left-exact contravariant functor. | |
\end{proposition} | |
We write this proof in slightly more detail than \cref{homcovleftexact}, | |
because of the | |
contravariance. | |
\begin{proof} | |
We want to show that $\hom(\cdot, M)$ is a left-exact contravariant functor, | |
which means that | |
if $ A \xrightarrow u B \xrightarrow v C \to 0$ is exact, then so is | |
$$ | |
0 \to \hom(C, M) \xrightarrow{\ol v} \hom(B, M) \xrightarrow{\ol u} \hom(A, M) | |
$$ | |
is exact. Here, the bold notation refers to the induced maps of $u,v$ on the | |
hom-sets: if $f \in \hom(B,M)$ and $g \in \hom(C, M)$, we define | |
$\ol u$ and $\ol v$ via $\ol v(g) = g \circ v$ and | |
$\ol u(f) = f \circ u$. | |
Let us show first that $\ol v$ is injective. | |
Suppose that $g \in \hom(C, M)$. If $\ol v(g) = g \circ v = 0$ then | |
$(g \circ v)(b) = 0$ for all $b \in B$. Since $v$ is a surjection, this means | |
that $g(C) = 0$ and hence $g = 0$. Therefore, $\ol v$ is injective, and we | |
have exactness at $\hom(C, M)$. | |
Since $v \circ u = 0$, it is clear that $\ol u \circ \ol u = 0$. | |
Now, suppose that $f \in \ker(\ol u) \subset \hom(B, M)$. Then | |
$\ol u(f) = f \circ u = 0$. | |
Thus $f: B \to M$ factors through $B/\im(u)$. | |
However, $\im(u) = \ker(v)$, so $f$ factors through $B/\ker(v)$. | |
Exactness shows that there is an isomorphism $B/\ker(v) \simeq C$. | |
In particular, we find that $f$ factors through $C$. This is what we wanted. | |
\end{proof} | |
\begin{exercise} | |
Come up with an example where $\hom_R(-, M)$ is not exact. | |
\end{exercise} | |
\begin{exercise} | |
Over a \emph{field}, $\hom$ is always exact. | |
\end{exercise} | |
\subsection{Projective modules} | |
Let $M$ be an $R$-module for a fixed commutative ring $R$. We have seen that | |
$\hom_R(M,-)$ is generally only a left-exact functor. | |
Sometimes, however, we do have exactness. We axiomatize this with the | |
following. | |
\begin{definition} \label{projectives} | |
An $R$-module $M$ is called \textbf{projective} if the functor $\hom_R(M, | |
\cdot)$ is | |
exact.\footnote{It is possible to define a projective module over a | |
noncommutative ring. The definition is the same, except that the $\hom$-sets | |
are no longer modules, but simply abelian groups. } | |
\end{definition} | |
One may first observe that a free module is projective. | |
Indeed, let $F = R^I$ for an indexing set. Then the functor $N \to \hom_R(F, | |
N)$ is | |
naturally | |
isomorphic to $N \to N^I$. It is easy to see that this functor preserves | |
exact sequences (that is, if $0 \to A \to B \to C \to 0$ is exact, so is $0 | |
\to A^I \to B^I \to C^I \to 0$). | |
Thus $F$ is projective. | |
One can also easily check that a \emph{direct summand} of a projective module | |
is projective. | |
It turns out that projective modules have a very clean characterization. They | |
are \emph{precisely} the direct | |
summands in free modules. | |
\add{check this} | |
\begin{proposition} \label{projmod} | |
The following are equivalent for an $R$-module $M$: | |
\begin{enumerate} | |
\item $M$ is projective. | |
\item Given any map $M \to N/N'$ from $M$ into a quotient of $R$-module | |
$N/N'$, we can lift | |
it to a map $M \to N$. | |
\item There is a module $M'$ such that $M \oplus M'$ is free. | |
\end{enumerate} | |
\end{proposition} | |
\begin{proof} | |
The equivalence of 1 and 2 is just unwinding the definition of projectivity, | |
because we just need to show that $\hom_R(M, \cdot)$ preserves surjective | |
maps, i.e. quotients. ($\hom_R(M, \cdot)$ is already left-exact, after all.) | |
To say that $\hom_R(M, N) \to \hom_R(M, N/N')$ is surjective is just the | |
statement that any map $M \to N/N'$ | |
can be lifted to $M \to N$. | |
Let us show that 2 implies 3. Suppose $M$ satisfies 2. Then choose a | |
surjection $P \twoheadrightarrow M$ where $P$ is free, by | |
\cref{freesurjection}. Then we can | |
write $M \simeq P/P'$ for a submodule $P' \subset P$. The isomorphism map | |
$M \to P/P'$ | |
leads by 2 to a lifting $M \to P$. In particular, there is a section of $P | |
\to M$, | |
namely this lifting. Since a section leads to a split exact sequence by | |
\cref{}, we find then that $P \simeq \ker(P \to M) \oplus \im(M \to P) \simeq | |
\ker(P \to M) \oplus M$, | |
verifying 3 since $P$ is free. | |
Now let us show that 3 implies 2. | |
Suppose $M \oplus M'$ is free, isomorphic to $P$. Then a map $M \to N/N'$ can | |
be extended to | |
\[ P \to N/N' \] | |
by declaring it to be trivial on $M'$. But now $P \to N/N'$ can be lifted to | |
$N$ because $P$ is free, and we have observed that a free module is | |
projective above; alternatively, we just lift the image of a basis. This | |
defines $P | |
\to N$. We may then compose this with the inclusion $M \to P$ to get the | |
desired map $M \to P \to N$, | |
which is a lifting of $M \to N/N'$. | |
\end{proof} | |
Of course, the lifting $P \to N$ of a given map $P \to N/N'$ is generally not | |
unique, and in fact is unique precisely when $\hom_R(P,N') = 0$. | |
So projective modules are precisely those with the following lifting property. | |
Consider a diagram | |
\[ \xymatrix{ | |
& P \ar[d] \\ | |
M \ar[r] & M'' \ar[r] & 0 | |
}\] | |
where the bottom row is exact. Then, if $P$ is projective, there is a lifting | |
$P \to M$ making commutative the diagram | |
\[ \xymatrix{ | |
& P \ar[d]\ar@{-->}[ld] \\ | |
M \ar[r] & M'' \ar[r] & 0 | |
}\] | |
\begin{corollary} | |
Let $M$ be a module. Then there is a surjection $P \twoheadrightarrow M$, | |
where $P$ is projective. | |
\end{corollary} | |
\begin{proof} | |
Indeed, we know (\rref{freesurjection}) that we can always get a surjection | |
from a free | |
module. Since free modules are projective by \rref{projmod}, we are | |
done. | |
\end{proof} | |
\begin{exercise} | |
Let $R$ be a principal ideal domain, $F'$ a submodule of a free module | |
$F$. Show that | |
$F'$ is free. (Hint: well-order the set of generators of $F$, and climb up by | |
transfinite induction.) | |
In particular, any projective modules is free. | |
\end{exercise} | |
\subsection{Example: the Serre-Swan theorem} | |
We now briefly digress to describe an important correspondence between | |
projective modules and vector bundles. The material in this section will not | |
be used in the sequel. | |
Let $X$ be a compact space. We shall not recall the topological notion of a | |
\emph{vector bundle} here. | |
We note, however, that if $E$ is a (complex) vector bundle, | |
then the set $\Gamma(X, E)$ of global sections is naturally a module over the | |
ring $C(X)$ of complex-valued continuous functions on $X$. | |
\begin{proposition} | |
If $E$ is a vector bundle on a compact Hausdorff space $X$, then there is a | |
surjection $\mathcal{O}^N \twoheadrightarrow E$ for some $N$. | |
\end{proposition} | |
Here $\mathcal{O}^N$ denotes the trivial bundle. | |
It is known that in the category of vector bundles, every epimorphism splits. | |
In particular, it follows that $E$ can be viewed as a \emph{direct summand} of | |
the bundle $\mathcal{O}^N$. Since $\Gamma(X, E)$ is then a direct summand of | |
$\Gamma(X, \mathcal{O}^N) = C(X)^N$, we find that $\Gamma(X, E)$ is a direct | |
summand of a projective $C(X)$-module. Thus: | |
\begin{proposition} | |
$\Gamma(X, E)$ is a finitely generated projective $C(X)$-module. | |
\end{proposition} | |
\begin{theorem}[Serre-Swan] | |
The functor $E \mapsto \Gamma(X, E)$ induces an equivalence of categories | |
between vector bundles on $X$ and finitely generated projective modules over | |
$C(X)$. | |
\end{theorem} | |
\subsection{Injective modules} | |
\label{ssecinj} | |
We have given a complete answer to the question of when the functor | |
$\hom_R(M,-)$ is exact. We have shown that there are a lot of such | |
\emph{projective} modules in the category of $R$-modules, enough that any | |
module admits a surjection from one such. | |
However, we now have to answer the dual question: when is the functor | |
$\hom_R(-, Q)$ exact? | |
Let us make the dual definition: | |
\begin{definition} | |
An $R$-module $Q$ is \textbf{injective} if the functor $\hom_R(-,Q)$ is exact. | |
\end{definition} | |
Thus, a module $Q$ over a ring $R$ is injective if | |
whenever $M \to N$ is an injection, and one has a map $M \to Q$, it can be | |
extended to $N \to Q$: in other words, $\hom_R(N,Q ) \to \hom_R(M,Q)$ is | |
surjective. | |
We can visualize this by a diagram | |
\[ \xymatrix{ | |
0 \ar[r] & M \ar[r] \ar[d] & N \ar@{-->}[ld] \\ | |
& Q | |
}\] | |
where the dotted arrow always exists if $Q$ is injective. | |
The notion is dual to projectivity, in some sense, so just as every module $M$ | |
admits an epimorphic map $P \to M$ for $P$ projective, we expect by duality | |
that every module admits a monomorphic map $M \to Q$ for $Q$ injective. | |
This is in fact true, but will require some work. | |
We start, first, with a fact about injective abelian groups. | |
\begin{theorem}\label{divisibleimpliesinj} | |
A divisible abelian group (i.e. one where the map $x | |
\to nx$ for any $n \in \mathbb{N}$ is surjective) is injective as a | |
$\mathbb{Z}$-module (i.e. abelian group). | |
\end{theorem} | |
\begin{proof} | |
The actual idea of the proof is rather simple, and similar to the proof | |
of the Hahn-Banach theorem. | |
Namely, we extend bit by bit, and then use Zorn's lemma. | |
The first step is that we have a subgroup $M $ of a larger abelian | |
group $N$. | |
We have a map of $f:M \to Q$ for $Q$ some divisible abelian group, and we | |
want to extend it to $N$. | |
Now we can consider the poset of pairs $(\tilde{f}, M')$ where $M' \supset | |
M$, and $\tilde{f}: M' \to N$ is a map extending $f$. | |
Naturally, we make this into a poset by defining the order as ``$(\tilde{f}, | |
M') \leq (\tilde{f}', M'')$ if $M'' $ contains $M'$ and $\tilde{f}'$ | |
is an extension of $\tilde{f}$. | |
It is clear that every chain has an upper bound, so Zorn's lemma implies | |
that we have a submodule $M' \subset N$ containing $M$, and a map $\tilde{f}: M' | |
\to N$ extending $f$, such that there is no proper extension of $\tilde{f}$. | |
From this we will derive a contradiction unless $M' = N$. | |
So suppose we have $M' \neq N$, for $M'$ the maximal submodule to which $f$ | |
can be extended, as in the above paragraph. Pick $m \in N - M'$, and consider the | |
submodule $M' + \mathbb{Z} m \subset N$. We are going to show how to extend $\tilde{f}$ | |
to this bigger submodule. First, suppose $\mathbb{Z}m \cap M' = \{0\}$, | |
i.e. the sum is direct. Then we can extend $\tilde{f}$ because $M' + | |
\mathbb{Z}m$ is a | |
direct sum: just define it to be zero on $\mathbb{Z}m$. | |
The slightly harder part is what happens if $\mathbb{Z} m \cap M' \neq \{ 0\}$. | |
In this case, there is an ideal $I \subset \mathbb{Z}$ such that $n \in I$ | |
if and only if $nm \in M'$. | |
This ideal, however, is principal; let $g \in \mathbb{Z} - \left\{0\right\}$ be a generator. Then $gm = p | |
\in M'$. In particular, $\tilde{f}(gm)$ is defined. | |
We can ``divide'' this | |
by $g$, i.e. find $u \in Q$ such that $gu = \tilde{f}(gm)$. | |
Now we may extend to a | |
map $\tilde{f}'$ from $\mathbb{Z} m + M'$ into $Q$ as follows. Choose $m' | |
\in M', k \in \mathbb{Z}$. Define $\tilde{f}'( m' + km) = \tilde{f}(m') | |
+ k u$. It is easy to see that this is well-defined by the choice of $u$, | |
and gives a proper extension of $\tilde{f}$. This contradicts maximality of | |
$M'$ and completes the proof. | |
\end{proof} | |
\begin{exercise} | |
\cref{divisibleimpliesinj} works over any principal ideal domain. | |
\end{exercise} | |
\begin{exercise}[Baer] \label{baercriterion} | |
Let $N$ be an $R$-module such that for any ideal $I \subset R$, any morphism | |
$I \to N$ can be extended to $R \to N$. Then $N$ is injective. (Imitate the | |
above argument.) | |
\end{exercise} | |
From this, we may prove: | |
\begin{theorem} | |
Any $R$-module $M$ can be imbedded in an injective $R$-module $Q$. | |
\end{theorem} | |
\begin{proof} | |
First of all, we know that any $R$-module $M$ is a quotient of a free | |
$R$-module. We are going to show that the {dual} (to be defined shortly) of a free module is injective. And so since | |
every module admits a surjection from a free module, we will use a dualization | |
argument to prove the present theorem. | |
First, for any abelian group $G$, define the \textbf{dual group} as $G^\vee | |
= \hom_{\mathbb{Z}}(G, \mathbb{Q}/\mathbb{Z})$. | |
Dualization is clearly a contravariant functor from abelian groups to abelian | |
groups. | |
By \cref{homcontleftexact} | |
and \cref{divisibleimpliesinj}, an exact | |
sequence of groups | |
\[ 0 \to A \to B \to C \to 0 \] | |
induces an exact sequence | |
\[ 0 \to C^\vee \to B^\vee \to A^\vee \to 0 .\] | |
In particular, dualization is an exact functor: | |
\begin{proposition} Dualization preserves exact sequences (but reverses | |
the order). | |
\end{proposition} | |
Now, we are going to apply this to $R$-modules. The dual of a left $R$-module | |
is acted upon by $R$. | |
The action, which is natural enough, is as follows. Let $M$ be an | |
$R$-module, and $f: M \to | |
\mathbb{Q}/\mathbb{Z}$ be a homomorphism of abelian groups (since | |
$\mathbb{Q}/\mathbb{Z}$ has in general no $R$-module structure), and $r \in | |
R$; then we define $rf$ to be the map $M \to \mathbb{Q}/\mathbb{Z}$ defined via | |
\[ (rf)(m) = f(rm).\] | |
It is easy to check that $M^{\vee}$ is thus made into an | |
$R$-module.\footnote{If $R$ is noncommutative, this would not work: instead | |
$M^{\vee}$ would be an \emph{right} $R$-module. For commutative rings, we have | |
no such distinction between left and right modules.} | |
In particular, dualization into $\mathbb{Q}/\mathbb{Z}$ gives a contravariant | |
exact functor from $R$-\emph{modules} to $R$-\emph{modules}. | |
Let $M$ be as before, | |
and now consider the $R$-module $M^{\vee}$. By \cref{freesurjection}, we can | |
find a free | |
module $F$ and a surjection | |
\[ F \to M^{\vee} \to 0.\] | |
Now dualizing gives an exact sequence of $R$-modules | |
\[ 0 \to M^{\vee \vee} \to F^{\vee}. \] | |
However, there is a natural map (of $R$-modules) $M \to M^{\vee \vee}$: given $m \in M$, we can | |
define a functional $\hom(M, \mathbb{Q}/\mathbb{Z}) \to \mathbb{Q}/\mathbb{Z}$ | |
by evaluation at $m$. One can check that this is a homomorphism. Moreover, this morphism $M \to M^{\vee \vee}$ is actually injective: if $m \in M$ were | |
in the kernel, then by definition every functional $M \to | |
\mathbb{Q}/\mathbb{Z}$ must vanish on $m$. It is easy to see (using | |
$\mathbb{Z}$-injectivity of $\mathbb{Q}/\mathbb{Z}$) that this cannot happen | |
if $m \neq 0$: we could just pick a nontrivial functional on the monogenic | |
\emph{subgroup} $\mathbb{Z} m$ and extend to $M$. | |
We claim now that $F^{\vee}$ is injective. This will prove the theorem, as | |
we have the composite of monomorphisms $M \hookrightarrow M^{\vee \vee} \hookrightarrow F^{\vee}$ that | |
embeds $M$ inside an injective module. | |
\begin{lemma} The dual of a free $R$-module $F$ is an injective | |
$R$-module. | |
\end{lemma} | |
\begin{proof} | |
Let $0 \to A \to B $ be exact; we have to show that | |
\[ \hom_R( B, F^\vee) \to \hom_R(A, F^\vee) \to 0 .\] | |
is exact. | |
Now we can reduce to the case where $F$ is the $R$-module $R$ itself. | |
Indeed, $F$ is a direct sum of $R$'s by assumption, and taking hom's turns | |
them into direct products; moreover the direct product of | |
exact sequences is exact. | |
So we are reduced to showing that $R^{\vee}$ is injective. | |
Now we claim that | |
\begin{equation} \label{weirddualityexpr} \hom_R(B, R^{\vee}) = | |
\hom_{\mathbb{Z}}(B, \mathbb{Q}/\mathbb{Z}). \end{equation} | |
In particular, $\hom_R( -, R^\vee)$ is an exact functor because | |
$\mathbb{Q}/\mathbb{Z}$ is an injective abelian group. | |
The proof of \cref{weirddualityexpr} is actually ``trivial.'' For instance, | |
a $R$-homomorphism $f: B \to R^\vee$ induces $\tilde{f}: B \to | |
\mathbb{Q}/\mathbb{Z}$ by sending $b \to (f(b))(1)$. One checks that this | |
is bijective. | |
\end{proof} | |
\end{proof} | |
\subsection{The small object argument} | |
There is another, more set-theoretic approach to showing that any $R$-module | |
$M$ can be imbedded in an injective module. | |
This approach, which constructs the injective module by a transfinite | |
colimit of push-outs, is essentially analogous to the ``small object | |
argument'' that one uses in homotopy theory to show that certain categories | |
(e.g. the category of CW complexes) are model categories in the sense of | |
Quillen; see \cite{Ho07}. | |
While this method is somewhat abstract and more complicated than the one of | |
\cref{ssecinj}, it is also more general. Apparently this method originates with Baer, | |
and was revisited by Cartan and Eilenberg in | |
\cite{Cartan-Eilenberg} and by Grothendieck in \cite{Gr57}. | |
There Grothendieck uses it to show that | |
many other abelian categories have enough injectives. | |
We first begin with a few remarks on smallness. | |
Let $\{B_{\alpha}\}, \alpha \in \mathcal{A}$ be an inductive system of objects in some | |
category $\mathcal{C}$, indexed by | |
an ordinal $\mathcal{A}$. Let us assume that $\mathcal{C}$ has (small) | |
colimits. If $A$ is an object of $\mathcal{C}$, then there is a | |
natural map | |
\begin{equation} \label{naturalmapcolim} \varinjlim \hom(A, B_\alpha) \to | |
\hom(A, \varinjlim B_\alpha) \end{equation} | |
because if one is given a map $A \to B_\beta$ for some $\beta$, one | |
naturally gets a map from $A$ into the colimit by composing with $B_\beta | |
\to \varinjlim B_\alpha$. (Note that the left colimit is one of sets!) | |
In general, the map \cref{naturalmapcolim} is neither injective or surjective. | |
\begin{example} | |
Consider the category of sets. Let $A = \mathbb{N}$ and $B_n = \left\{1, | |
\dots, n\right\}$ be the inductive system indexed by the natural numbers | |
(where $B_n \to B_{m}, n \leq m$ is the obvious map). Then $\varinjlim B_n = | |
\mathbb{N}$, so there is a map | |
\[ A \to \varinjlim B_n, \] | |
which does not factor as | |
\[ A \to B_m \] | |
for any $m$. Consequently, $\varinjlim \hom(A, B_n) \to \hom(A, \varinjlim | |
B_n)$ is not surjective. | |
\end{example} | |
\begin{example} | |
Next we give an example where the map fails to be injective. Let $B_n = | |
\mathbb{N}/\left\{1, 2, \dots, n\right\}$, that is, the quotient set of | |
$\mathbb{N}$ with the first $n$ elements collapsed to one element. | |
There are natural maps $B_n \to B_m$ for $n \leq m$, so the | |
$\left\{B_n\right\}$ form an inductive system. It is easy to see that the | |
colimit $\varinjlim B_n = \left\{\ast \right\}$: it is the one-point set. | |
So it follows that $\hom(A, \varinjlim B_n)$ is a one-element set. | |
However, $\varinjlim \hom(A , B_n)$ is \emph{not} a one-element set. | |
Consider the family of maps $A \to B_n$ which are just the natural projections | |
$\mathbb{N} \to \mathbb{N}/\left\{1, 2, \dots, n\right\}$ and the family of | |
maps $A \to B_n$ which map the whole of $A$ to the class of $1$. | |
These two families of maps are distinct at each step and thus are distinct in | |
$\varinjlim \hom(A, B_n)$, but they induce the same map $A \to \varinjlim B_n$. | |
\end{example} | |
Nonetheless, if $A$ is a \emph{finite set}, it is easy to see that for any | |
sequence of sets $B_1 \to B_2 \to \dots$, we have | |
\[ \varinjlim \hom(A, B_n) = \hom(A, \varinjlim B_n). \] | |
\begin{proof} | |
Let $f: A \to \varinjlim B_n$. The range of $A$ is finite, containing say | |
elements $c_1, \dots, c_r \in \varinjlim B_n$. These all come from some | |
elements in $B_N$ for $N$ large by definition of the colimit. Thus we can | |
define $\widetilde{f}: A \to B_N$ lifting $f$ at a finite stage. | |
Next, suppose two maps $f_n: A \to B_m, | |
g_n : A \to B_m$ define the same map $A \to \varinjlim B_n$. | |
Then each of the finitely many elements of $A$ gets sent to the same point in | |
the colimit. By definition of the colimit for sets, there is $N \geq m$ such | |
that the finitely many elements of $A$ get sent to the same points in $B_N$ | |
under $f$ and $g$. This shows that $\varinjlim \hom(A, B_n) \to \hom(A, | |
\varinjlim B_n)$ is injective. | |
\end{proof} | |
The essential idea is that $A$ is ``small'' relative to the long chain of | |
compositions $B_1 \to B_2 \to \dots$, so that it has to factor through a | |
finite step. | |
Let us generalize this. | |
\begin{definition} \label{smallness} | |
Let $\mathcal{C}$ be a category, $I $ a class of maps, and $\omega$ an ordinal. | |
An object $A \in \mathcal{C}$ is said to be $\omega$-\textbf{small} (with | |
respect to $I$) if | |
whenever $\{B_\alpha\}$ is an inductive system parametrized by $\omega$ with | |
maps in $I$, then | |
the map | |
\[ \varinjlim \hom(A, B_\alpha) \to \hom(A, \varinjlim B_\alpha) \] | |
is an isomorphism. | |
\end{definition} | |
Our definition varies slightly from that of \cite{Ho07}, where only ``nice'' | |
transfinite sequences $\left\{B_\alpha\right\}$ are considered. | |
In our applications, we shall begin by restricting ourselves to the category | |
of $R$-modules for a fixed commutative ring $R$. | |
We shall also take $I$ to be the set of \emph{monomorphisms,} or | |
injections.\footnote{There are, incidentally, categories, such as the category | |
of rings, where a categorical epimorphism may not be a surjection of sets.} | |
Then each of the maps | |
\[ B_\beta \to \varinjlim B_\alpha \] | |
is an injection, so it follows that | |
$\hom(A, B_\beta )\to \hom (A, \varinjlim B_\alpha)$ is one, and in | |
particular the canonical map | |
\begin{equation} \label{homcolimmap} \varinjlim \hom(A, B_\alpha) \to \hom (A, | |
\varinjlim B_\alpha) \end{equation} | |
is an \emph{injection.} | |
We can in fact interpret the $B_\alpha$'s as subobjects of the big module | |
$\varinjlim B_\alpha$, and think of their union as $\varinjlim B_\alpha$. | |
(This is not an abuse of notation if we identify $B_\alpha$ with the image in | |
the colimit.) | |
We now want to show that modules are always small for ``large'' ordinals | |
$\omega$. | |
For this, we have to digress to do some set theory: | |
\begin{definition} | |
Let $\omega$ be a \emph{limit} ordinal, and $\kappa$ a cardinal. Then $\omega$ is | |
\textbf{$\kappa$-filtered} if every collection $C$ of ordinals strictly less | |
than $\omega$ and of cardinality at most $\kappa$ has an upper bound strictly | |
less than $\omega$. | |
\end{definition} | |
\begin{example} \label{limitordfinfiltered} | |
A limit ordinal (e.g. the natural numbers $\omega_0$) is $\kappa$-filtered for any finite cardinal $\kappa$. | |
\end{example} | |
\begin{proposition} | |
Let $\kappa$ be a cardinal. Then there exists a $\kappa$-filtered ordinal | |
$\omega$. | |
\end{proposition} | |
\begin{proof} | |
If $\kappa$ is finite, \cref{limitordfinfiltered} shows that any limit ordinal | |
will do. Let us thus assume that $\kappa$ is infinite. | |
Consider the smallest ordinal $\omega$ whose cardinality is strictly greater | |
than that of $\kappa$. Then we claim that $\omega$ is $\kappa$-filtered. | |
Indeed, if $C$ is a collection of at most $\kappa$ ordinals strictly smaller | |
than $\omega$, then each of these ordinals is of size at most $\kappa$. Thus | |
the union of all the ordinals in $C$ (which is an ordinal) is of size at most | |
$\kappa$, so is strictly smaller than $\omega$, and it provides an upper bound as in the definition. | |
\end{proof} | |
\begin{proposition} \label{modulesaresmall} | |
Let $M$ be a module, $\kappa$ the cardinality of the set of its submodules. | |
Then if $\omega$ is $\kappa$-filtered, then $M$ is $\omega$-small (with | |
respect to injections). | |
\end{proposition} | |
The proof is straightforward, but let us first think about a special case. If | |
$M$ is finite, then the claim is that for any inductive system | |
$\left\{B_\alpha\right\}$ with injections between them, parametrized by a | |
limit ordinal, any map $M \to | |
\varinjlim B_\alpha$ factors through one of the $B_\alpha$. But this is clear. | |
$M$ is finite, so since each element in the image must land inside one of the | |
$B_\alpha$, so all of $M$ lands inside some finite stage. | |
\begin{proof} | |
We need only show that the map \cref{homcolimmap} is a surjection when | |
$\omega$ is $\kappa$-filtered. | |
Let $f: A \to \varinjlim B_\alpha$ be a map. | |
Consider the subobjects $\{f^{-1}(B_\alpha)\}$ of $A$, where $B_\alpha$ is considered as a | |
subobject of the colimit. If one of these, say $f^{-1}(B_\beta)$, fills $A$, | |
then the map factors through $B_\beta$. | |
So suppose to the contrary that all of the $f^{-1}(B_\alpha)$ were proper | |
subobjects of $A$. | |
However, we know that | |
\[ \bigcup f^{-1}(B_\alpha) = f^{-1}\left(\bigcup B_\alpha\right) = A. \] | |
Now there are at most $\kappa$ different subobjects of $A$ that occur among | |
the $f^{-1}(B_\alpha)$, by hypothesis. | |
Thus we can find a set $A$ of cardinality at most $\kappa$ such that as | |
$\alpha'$ ranges over $A$, the | |
$f^{-1}(B_{\alpha'})$ range over \emph{all} the $f^{-1}(B_\alpha)$. | |
However, $A$ has an upper bound $\widetilde{\omega} < \omega$ as $\omega$ is | |
$\kappa$-filtered. In particular, | |
all the $f^{-1}(B_{\alpha'})$ are contained in | |
$f^{-1}(B_{\widetilde{\omega}})$. It follows that | |
$f^{-1}(B_{\widetilde{\omega}}) = A$. | |
In particular, the map $f$ factors through $B_{\widetilde{\omega}}$. | |
\end{proof} | |
From this, we will be able to deduce the existence of lots of injectives. | |
Let us recall the criterion of Baer (\cref{baercriterion}): a module $Q$ is | |
injective if and only if in every commutative diagram | |
\[ \xymatrix{ | |
\mathfrak{a} \ar[d] \ar[r] & Q \\ | |
R \ar@{-->}[ru] | |
}\] | |
for $\mathfrak{a} \subset R$ an ideal, the dotted arrow exists. In other | |
words, we are trying to solve an \emph{extension problem} with respect to the | |
inclusion $\mathfrak{a} \hookrightarrow R$ into the module $M$. | |
If $M$ is an $R$-module, then in general we may have a semi-complete diagram as above. In | |
it, we can form the \emph{push-out} | |
\[ \xymatrix{ | |
\mathfrak{a} \ar[d] \ar[r] & Q \ar[d] \\ | |
R \ar[r] & R \oplus_{\mathfrak{a}} Q | |
}.\] | |
Here the vertical map is injective, and the diagram commutes. The point is | |
that we can extend $\mathfrak{a} \to Q$ to $R$ \emph{if} we extend $Q$ to the | |
larger module $R \oplus_{\mathfrak{a}} Q$. | |
The point of the small object argument is to repeat this procedure | |
transfinitely many times. | |
\begin{theorem} | |
Let $M$ be an $R$-module. Then there is an embedding $M \hookrightarrow Q$ for | |
$Q$ injective. | |
\end{theorem} | |
\begin{proof} | |
We start by defining a functor $\mathbf{M}$ on the category of $R$-modules. | |
Given $N$, we consider the set of all maps $\mathfrak{a} \to N$ for | |
$\mathfrak{a} \subset R$ an ideal, and consider the push-out | |
\begin{equation} \label{hugediag} | |
\xymatrix{ | |
\bigoplus \mathfrak{a}\ar[r] \ar[d] & N \ar[d] \\ | |
\bigoplus R \ar[r] & N \oplus_{\bigoplus \mathfrak{a}} \bigoplus R | |
} | |
\end{equation} | |
where the direct sum of copies of $R$ is taken such that every copy of an | |
ideal $\mathfrak{a}$ corresponds to one copy of $R$. | |
We define $\mathbf{M}(N)$ to be this push-out. Given a map $N \to N'$, there | |
is a natural morphism of diagrams \cref{hugediag}, so $\mathbf{M}$ is a | |
functor. | |
Note furthermore that there is a natural transformation | |
\[ N \to \mathbf{M}(N), \] | |
which is \emph{always an injection.} | |
The key property of $\mathbf{M}$ is that if $\mathfrak{a} \to N$ is any | |
morphism, it can be extended to $R \to \mathbf{M}(N)$, by the very | |
construction of $\mathbf{M}(N)$. The idea will now be to | |
apply $\mathbf{M}$ a transfinite number of times and to use the small object | |
property. | |
We define for each ordinal $\omega$ a functor $\mathbf{M}_{\omega}$ on the | |
category of $R$-modules, together with a natural injection $N \to | |
\mathbf{M}_{\omega}(N)$. We do this by transfinite induction. | |
First, $\mathbf{M}_1 = \mathbf{M}$ is the functor defined above. | |
Now, suppose given an ordinal $\omega$, and suppose $\mathbf{M}_{\omega'}$ is | |
defined for $\omega' < \omega$. If $\omega$ has an immediate predecessor | |
$\widetilde{\omega}$, we let | |
$$\mathbf{M}_{\omega} = \mathbf{M} \circ \mathbf{M}_{\widetilde{\omega}}.$$ | |
If not, we let $\mathbf{M}_{\omega}(N) = \varinjlim_{\omega' < \omega} | |
\mathbf{M}_{\omega'}(N)$. | |
It is clear (e.g. inductively) that the $\mathbf{M}_{\omega}(N)$ form an inductive system over | |
ordinals $\omega$, so this is reasonable. | |
Let $\kappa$ be the cardinality of the set of ideals in $R$, and let $\Omega$ | |
be a $\kappa$-filtered ordinal. | |
The claim is as follows. | |
\begin{lemma} | |
For any $N$, $\mathbf{M}_{\Omega}(N)$ is injective. | |
\end{lemma} | |
If we prove this, we will be done. In fact, we will have shown that there is a | |
\emph{functorial} embedding of a module into an injective. | |
Thus, we have only to prove this lemma. | |
\begin{proof} | |
By Baer's criterion (\cref{baercriterion}), it suffices to show that if | |
$\mathfrak{a} \subset R$ is an ideal, then any map $f: \mathfrak{a} \to | |
\mathbf{M}_{\Omega}(N)$ extends to $R \to \mathbf{M}_{\Omega}(N)$. However, we | |
know since $\Omega$ is a limit ordinal that | |
\[ \mathbf{M}_{\Omega}(N) = \varinjlim_{\omega < \Omega} | |
\mathbf{M}_{\omega}(N), \] | |
so by \cref{modulesaresmall}, we find that | |
\[ \hom_R(\mathfrak{a}, \mathbf{M}_{\Omega}(N)) = \varinjlim_{\omega < \Omega} | |
\hom_R(\mathfrak{a}, \mathbf{M}_{\omega}(N)). \] | |
This means in particular that there is some $\omega' < \Omega$ such that $f$ | |
factors through the submodule $\mathbf{M}_{\omega'}(N)$, as | |
\[ f: \mathfrak{a} \to \mathbf{M}_{\omega'}(N) \to \mathbf{M}_{\Omega}(N). \] | |
However, by the fundamental property of the functor $\mathbf{M}$, we know that | |
the map $\mathfrak{a} \to \mathbf{M}_{\omega'}(N)$ can be extended to | |
\[ R \to \mathbf{M}( \mathbf{M}_{\omega'}(N)) = \mathbf{M}_{\omega' + 1}(N), \] | |
and the last object imbeds in $\mathbf{M}_{\Omega}(N)$. | |
In particular, $f$ can be extended to $\mathbf{M}_{\Omega}(N)$. | |
\end{proof} | |
\end{proof} | |
\subsection{Split exact sequences} | |
\add{additive functors preserve split exact seq} | |
Suppose that | |
$\xymatrix@1{0 \ar[r] & L \ar[r]^\psi & M \ar[r]^f & N \ar[r] & 0}$ | |
is a split short exact sequence. | |
Since $\Hom_R (D, \cdot)$ is a left-exact functor, we see that | |
$$\xymatrix@1{0 \ar[r] | |
& \Hom_R(D, L) \ar[r]^{\psi'} | |
& \Hom_R(D, M) \ar[r]^{\f'} | |
& \Hom_R(D, N)}$$ | |
is exact. In addition, | |
$\Hom_R (D, L \oplus N) \cong \Hom_R(D, L) \oplus \Hom_R (D, N)$. Therefore, in | |
the case that we start with a split short exact sequence $M \cong L \oplus N$, | |
applying $\Hom_R (D, \cdot)$ does yield a split short exact sequence | |
$$\xymatrix@1{0 \ar[r] | |
& \Hom_R(D, L) \ar[r]^{\psi'} | |
& \Hom_R(D, M) \ar[r]^{\f'} | |
& \Hom_R(D, N) \ar[r] & 0}.$$ | |
Now, assume that | |
$$\xymatrix@1{0 \ar[r] | |
& \Hom_R(D, L) \ar[r]^{\psi'} | |
& \Hom_R(D, M) \ar[r]^{\f'} | |
& \Hom_R(D, N) \ar[r] & 0}$$ | |
is a short exact sequence of abelian groups for all $R$-modules $D$. | |
Set $D = R$ and using $\Hom_R (R, N) \cong N$ yields that | |
$\xymatrix@1{0 \ar[r] & L \ar[r]^\psi & M \ar[r]^f & N \ar[r] & 0}$ | |
is a short exact sequence. | |
Set $D = N$, so we have | |
$$\xymatrix@1{0 \ar[r] | |
& \Hom_R(N, L) \ar[r]^{\psi'} | |
& \Hom_R(N, M) \ar[r]^{\f'} | |
& \Hom_R(N, N) \ar[r] & 0}$$ | |
Here, $\f'$ is surjective, so the identity map of $\Hom_R (N, N)$ lifts to a | |
map $g \in \Hom_R (N, M)$ so that $f \circ g = \f'(g) = id$. | |
This means that $g$ is a splitting homomorphism for the sequence | |
$\xymatrix@1{0 \ar[r] & L \ar[r]^\psi & M \ar[r]^f & N \ar[r] & 0}$, | |
and therefore the sequence is a split short exact sequence. | |
\section{The tensor product} | |
\label{sec:tensorprod} | |
We shall now introduce the third functor of this chapter: the tensor product. | |
The tensor product's key property is that it allows one to ``linearize'' | |
bilinear maps. When taking the tensor product of rings, it provides a | |
categorical coproduct as well. | |
\subsection{Bilinear maps and the tensor product} | |
Let $R$ be a commutative ring, as usual. | |
We have seen that the $\hom$-sets $\hom_R(M,N)$ of $R$-modules $M,N$ are themselves | |
$R$-modules. | |
Consequently, if we have three $R$-modules $M,N,P$, we can think about | |
module-homomorphisms | |
\[ M \stackrel{\lambda}{\to}\hom_R(N,P). \] | |
Suppose $x \in M, y \in N$. Then we can consider | |
\( \lambda(x) \in \hom_R(N,P) \) | |
and thus we can consider the element | |
\( \lambda(x)(y) \in P. \) | |
We denote this element $\lambda(x)(y)$, which depends on the variables $x \in | |
M, y \in N$, by $\lambda(x,y)$ for convenience; it | |
is a function of two variables $M \times N \to P$. | |
There are | |
certain properties of $\lambda(\cdot, \cdot)$ that we list below. | |
Fix $x , x' \in M$; $y, y' \in N; \ a \in R$. Then: | |
\begin{enumerate} | |
\item $\lambda(x,y+y') = \lambda(x,y) + \lambda(x, y')$ because $\lambda(x)$ | |
is | |
additive. | |
\item $\lambda(x, ay) = a \lambda(x,y)$ because $\lambda(x)$ is an | |
$R$-module homomorphism. | |
\item $\lambda(x+x', y) = \lambda(x,y) + \lambda(x', y)$ because | |
$\lambda$ is additive. | |
\item $\lambda(ax, y) = a\lambda(x,y)$ because $\lambda$ is an $R$-module | |
homomorphism. | |
\end{enumerate} | |
Conversely, given a function $\lambda: M \times N \to P$ of two variables satisfying the above properties, | |
it is easy to see that we can get a morphism of $R$-modules $M \to | |
\hom_R(N,P)$. | |
\begin{definition} | |
An \textbf{$R$-bilinear map $\lambda: M \times N \to P$} is a map satisfying | |
the above listed conditions. In other words, it is required to be $R$-linear | |
in each variable separately. | |
\end{definition} | |
The previous discussion shows that there is a \emph{bijection} between $R$-bilinear | |
maps $M \times N \to P$ with $R$-module maps $M \to \hom_R(N,P)$. | |
Note that the first interpretation is symmetric in $M,N$; the second, by | |
contrast, can be interpreted in terms of the old concepts of an $R$-module map. | |
So both are useful. | |
\begin{exercise} | |
Prove that a $\mathbb{Z}$-bilinear map out of $\mathbb{Z}/2 \times | |
\mathbb{Z}/3$ is identically zero, whatever the target module. | |
\end{exercise} | |
Let us keep the notation of the previous discussion: in particular, $M,N, P$ will | |
be modules over a commutative ring $R$. | |
Given a bilinear map $M \times N \to P$ and a homomorphism $P \to P'$, we can | |
clearly get a bilinear map $M \times N \to P'$ by composition. | |
In particular, given $M,N$, there is a \emph{covariant functor} from | |
$R$-modules to | |
$\mathbf{Sets}$ sending any $R$-module $P$ to the collection of $R$-bilinear | |
maps $M \times N | |
\to P$. As usual, we are interested in when this functor is | |
\emph{corepresentable.} | |
As a result, | |
we are interested in \emph{universal} bilinear maps out of $M \times N$. | |
\begin{definition} | |
An $R$-bilinear map $\lambda: M \times N \to P$ is called \textbf{universal} if | |
for all $R$-modules $Q$, the composition of $P \to Q$ with $M \times N | |
\stackrel{\lambda}{\to} P$ | |
gives a \textbf{bijection} | |
\[ \hom_R(P,Q) \simeq \left\{\mathrm{bilinear \ maps} \ M \times N \to | |
Q\right\} \] | |
So, given a bilinear map $M \times N \to Q$, there is a \textit{unique} map $P | |
\to Q$ making the diagram | |
\[ | |
\xymatrix{ | |
& P \ar[dd] \\ | |
M \times N \ar[ru]^{\lambda} \ar[rd] & \\ | |
& Q | |
} | |
\] | |
Alternatively, $P$ \emph{corepresents} the functor $Q \to | |
\left\{\mathrm{bilinear \ maps \ } M \times N \to Q\right\}$. | |
\end{definition} | |
General nonsense says that given $M,N$, an universal $R$-bilinear map $M | |
\times N \to P$ is | |
\textbf{unique} up to isomorphism (if it exists). This follows from \emph{Yoneda's lemma}. | |
For convenience, we give a direct proof. | |
Suppose $M \times N \stackrel{\lambda}{\to} P$ was universal and $M \times N | |
\stackrel{\lambda'}{\to} P'$ is also | |
universal. Then by the universal property, there are unique maps $P \to P'$ | |
and $P' \to P$ making the | |
following diagram commutative: | |
\[ | |
\xymatrix{ | |
& P \ar[dd] \\ | |
M \times N \ar[ru]^{\lambda} \ar[rd]^{\lambda'} & \\ | |
& P' \ar[uu] | |
} | |
\] | |
These compositions $P \to P' \to P, P' \to P \to P'$ have to be the identity | |
because of the uniqueness part of the universal property. | |
As a result, $P \to P'$ is an isomorphism. | |
We shall now show that this universal object does indeed exist. | |
\begin{proposition} \label{tensorexists} | |
Given $M,N$, a universal bilinear map out of $M \times N$ exists. | |
\end{proposition} | |
Before proving it we make: | |
\begin{definition} | |
We denote the codomain of the universal map out of $M \times N $ by $M | |
\otimes_R N$. This is called the \textbf{tensor product} of $M,N$, so there | |
is a universal bilinear map out of $M \times N$ into $M \otimes_R N$. | |
\end{definition} | |
\begin{proof}[Proof of \rref{tensorexists}] We will simply give | |
a presentation of the tensor product by | |
``generators and relations.'' | |
Take the free $R$-module $M \otimes_R N$ generated by the symbols $\left\{x | |
\otimes | |
y\right\}_{x \in M, y \in N}$ and quotient out by the relations forced upon us | |
by the definition of a bilinear map (for $x, x' \in M, \ y, y' \in N, \ a | |
\in R$) | |
\begin{enumerate} | |
\item $(x+x') \otimes y = x \otimes y + x' \otimes y$. | |
\item $(ax) \otimes y = a(x \otimes y) = x \otimes (ay)$. | |
\item $x \otimes (y+y') = x \otimes y + x \otimes y'$. | |
\end{enumerate} | |
We will abuse notation and denote $x \otimes y$ for its image in $M \otimes_R | |
N$ (as opposed to the symbol generating the free module). | |
There is a bilinear map $M \times N \to M \otimes_R N$ sending $(x,y) \to x | |
\otimes y$; the relations imposed imply that this map is a bilinear map. We | |
have to check | |
that it is universal, but this is actually quite direct. | |
Suppose we had a bilinear map $\lambda: M \times N \to P$. We must construct | |
a linear map $M | |
\otimes_R N \to P$. | |
To do this, we can just give a map on generators, and show that it is zero on | |
each of the relations. | |
It is easy to see that to make the appropriate diagrams commute, the linear | |
map $M \otimes N \to P$ has to send $x \otimes y \to \lambda(x,y)$. | |
This factors | |
through the relations on $x \otimes y$ by bilinearity and leads to an | |
$R$-linear map $M \otimes_{R} N \to P$ such that the following diagram | |
commutes: | |
\[ | |
\xymatrix{ | |
M \times N \ar[r] \ar[rd]^{\lambda} & M \otimes_R N \ar[d] \\ | |
& P | |
}.\] | |
It is easy to see that $M \otimes_R N \to P$ is unique because the $x \otimes | |
y$ generate it. | |
\end{proof} | |
The theory of the tensor product allows one to do away with bilinear maps and | |
just think of linear maps. | |
Given $M, N$, we have constructed an object $M \otimes_R N$. We now wish to see | |
the functoriality of the tensor product. In fact, $(M,N) \to M \otimes_R N$ is a \emph{covariant | |
functor} in two variables from $R$-modules to $R$-modules. | |
In particular, if $M \to M', N \to N'$ are morphisms, there is a canonical map | |
\begin{equation} \label{tensorisfunctor} M \otimes_R N \to M' \otimes_R N'. | |
\end{equation} | |
To obtain \cref{tensorisfunctor}, we take the natural bilinear map $M \times N \to M' \times N' | |
\to M' \otimes_R N'$ and use the universal property of $M \otimes_R N$ to get | |
a map out of it. | |
\subsection{Basic properties of the tensor product} | |
We make some observations and prove a few basic properties. As the proofs will | |
show, one powerful way to prove things about an object is to reason about its | |
universal property. If two objects have the same universal property, they are | |
isomorphic. | |
\begin{proposition} | |
The tensor product is symmetric: for $R$-modules $M,N$, we have $M \otimes_R | |
N \simeq N \otimes_R M$ | |
canonically. | |
\end{proposition} | |
\begin{proof} | |
This is clear from the universal properties: giving a bilinear map | |
out of $M \times N$ is the same as a bilinear map out $N \times M$. | |
Thus $M \otimes_R N$ and $N \otimes_R N$ have the same universal property. | |
It is also | |
clear from the explicit construction. | |
\end{proof} | |
\begin{proposition} | |
For an $R$-module $M$, there is a canonical isomorphism $M \to M \otimes_R R$. | |
\end{proposition} | |
\begin{proof} | |
If we think in terms of | |
bilinear maps, this statement is equivalent to the statement that a bilinear | |
map $\lambda: M \times R \to P$ is the same as a linear map $M \to N$. Indeed, | |
to do | |
this, restrict $\lambda$ to $\lambda(\cdot, 1)$. Given $f: M \to N$, | |
similarly, we take for $\lambda$ as $\lambda(x,a) = af(x)$. This gives a | |
bijection as claimed. | |
\end{proof} | |
\begin{proposition} | |
The tensor product is associative. There are canonical isomorphisms $M | |
\otimes_R (N \otimes_R P) \simeq (M | |
\otimes_R N) \otimes_R P$. | |
\end{proposition} | |
\begin{proof} | |
There are a few ways to see this: one is to build | |
it explicitly from the construction given, sending $x \otimes (y \otimes z) \to | |
(x \otimes y) \otimes z$. | |
More conceptually, both have the same universal | |
property: by general categorical nonsense (Yoneda's lemma), we need to show | |
that for all $Q$, there is a canonical bijection | |
\[ \hom_R(M \otimes (N \otimes P)), Q) \simeq \hom_R( (M \otimes N) | |
\otimes P, Q) \] | |
where the $R$'s are dropped for simplicity. But both of these sets can be | |
identified with the set of trilinear maps\footnote{Easy to define.} $M \times N | |
\times P \to Q$. Indeed | |
\begin{align*} | |
\hom_R(M \otimes (N \otimes P), Q) & \simeq \mathrm{bilinear} \ M \times (N | |
\otimes P) \to Q \\ | |
& \simeq \hom(N \otimes P, \hom(M,Q)) \\ | |
& \simeq \mathrm{bilinear} \ N \times P \to \hom(M,Q) \\ | |
& \simeq \hom(N, \hom(P, \hom(M,Q)) \\ | |
& \simeq \mathrm{trilinear\ maps}. | |
\end{align*} | |
\end{proof} | |
\subsection{The adjoint property} | |
Finally, while we defined the tensor product in terms of a ``universal | |
bilinear map,'' we saw earlier that bilinear maps could be interpreted as maps | |
into a suitable $\hom$-set. | |
In particular, fix $R$-modules $M,N,P$. We know that the set of bilinear maps | |
$M \times N \to P$ is naturally in bijection with | |
\[ \hom_R(M, \hom_R(N,P)) \] | |
as well as with | |
\[ \hom_R(M \otimes_R, N, P). \] | |
As a result, we find: | |
\begin{proposition} For $R$-modules $M,N,P$, there is a natural bijection | |
\[ \hom_R(M,\hom_R(N,P)) \simeq \hom_R(M \otimes_R N, P). \] | |
\end{proposition} | |
There is a more evocative way of phrasing the above natural bijection. Given | |
$N$, let us define the functors $F_N, G_N$ via | |
\[ F_N(M) = M \otimes_R N, \quad G_N(P) = \hom_R(N,P). \] | |
Then the above proposition states that there is a natural isomorphism | |
\[ \hom_R( F_N(M), P) \simeq \hom_R( M, G_N(P)). \] | |
In particular, $F_N$ and $G_N$ are \emph{adjoint functors}. So, in a sense, | |
the operations of $\hom$ and $\otimes$ are dual to each other. | |
\begin{proposition} \label{tensorcolimit} | |
Tensoring commutes with colimits. | |
\end{proposition} | |
In particular, it follows that if $\left\{N_\alpha\right\}$ is a family of | |
modules, and $M$ is a module, then | |
\[ M \otimes_R \bigoplus N_\alpha = \bigoplus M \otimes_R N_\alpha. \] | |
\begin{exercise} | |
Give an explicit proof of the above relation. | |
\end{exercise} | |
\begin{proof} | |
This is a formal consequence of the fact that the tensor product is a left | |
adjoint and consequently commutes with all colimits. | |
\add{proof} | |
\end{proof} | |
In particular, by \cref{tensorcolimit}, the tensor product commutes with \emph{cokernels.} | |
That is, if $A \to B \to C \to 0$ is an exact sequence of $R$-modules and $M$ | |
is an $R$-module, $A \otimes_R M \to B \otimes_R M \to C \otimes_R M \to 0$ is | |
also exact, because exactness of such a sequence is precisely a condition on | |
the cokernel. | |
That is, the tensor product is \emph{right exact.} | |
We can thus prove a simple result on finite generation: | |
\begin{proposition} \label{fingentensor} | |
If $M, N$ are finitely generated, then $M \otimes_R N$ is finitely generated. | |
\end{proposition} | |
\begin{proof} | |
Indeed, if we have surjections $R^m \to M, R^n \to N$, we can tensor them; we | |
get a surjection since the tensor product is right-exact. | |
So have a surjection | |
$R^{m n} = R^m \otimes_R R^n \to M \otimes_R N$. | |
\end{proof} | |
\subsection{The tensor product as base-change} | |
Before this, we have considered the tensor product as a functor within a | |
fixed category. Now, we shall see that when one takes the tensor product with a | |
\emph{ring}, one gets additional structure. As a result, we will be able to | |
get natural functors between \emph{different} module categories. | |
Suppose we have a | |
ring-homomorphism $\phi:R \to R'$. In this case, any $R'$-module can be | |
regarded as | |
an $R$-module. | |
In particular, there is a canonical functor of \emph{restriction} | |
\[ R'\mbox{-}\mathrm{modules} \to R\mbox{-}\mathrm{modules}. \] | |
We shall see that the tensor product provides an \emph{adjoint} to this | |
functor. | |
Namely, if $M$ has an $R$-module | |
structure, then $M \otimes_R R'$ has an $R'$ module structure where $R'$ acts | |
on the right. Since the tensor product is functorial, this gives a functor | |
in the opposite direction: | |
\[ R\mbox{-}\mathrm{modules} \to R'\mbox{-}\mathrm{modules}. \] | |
Let $M'$ be an $R'$-module and $M$ an $R$-module. In view of the above, | |
we can talk about | |
\[ \hom_R(M, M') \] | |
by thinking of $M'$ as an $R$-module. | |
\begin{proposition} | |
There is a canonical isomorphism between | |
\[ \hom_R(M, M') \simeq \hom_{R'}(M \otimes_R R', M'). \] | |
In particular, the restriction functor and the functor $M \to M \otimes_R R'$ | |
are adjoints to each other. | |
\end{proposition} | |
\begin{proof} | |
We can describe the bijection explicitly. Given an $R'$-homomorphism $f:M | |
\otimes_R R' \to M'$, we get a map | |
\[ f_0:M \to M' \] | |
sending | |
\[ m \to m \otimes 1 \to f(m \otimes 1). \] | |
This is easily seen to be an $R$-module-homomorphism. Indeed, | |
\[ f_0(ax) = f(ax \otimes 1) = f(\phi(a)(x \otimes 1)) = a f(x \otimes 1) = | |
a f_0(x) \] | |
since $f$ is an $R'$-module homomorphism. | |
Conversely, if we are given a homomorphism of $R$-modules | |
\[ f_0: M \to M' \] | |
then we can define | |
\[ f: M \otimes_R R' \to M' \] | |
by sending $m \otimes r' \to r' f_0(m)$, which is a homomorphism of $R'$ | |
modules. | |
This is well-defined because $f_0$ is a homomorphism of $R$-modules. We leave | |
some details to the reader. | |
\end{proof} | |
\begin{example} | |
In the representation theory of finite groups, the operation of tensor product | |
corresponds to the procedure of \emph{inducing} a representation. Namely, if | |
$H \subset G$ is a subgroup of a group $G$, then there is an obvious | |
restriction functor from $G$-representations to $H$-representations. | |
The adjoint to this is the induction operator. Since a $H$-representation | |
(resp. a $G$-representation) is just a module over the group ring, the | |
operation of induction is really a special case of the tensor product. Note | |
that the group rings are generally not commutative, so this should be | |
interpreted with some care. | |
\end{example} | |
\subsection{Some concrete examples} | |
We now present several concrete computations of tensor products in explicit | |
cases to illuminate what is happening. | |
\begin{example} Let us compute $\mathbb{Z}/10 \otimes_{\mathbb{Z}} | |
\mathbb{Z}/12$. | |
Since $1$ spans $\mathbb{Z} / (10)$ and $1$ spans $\mathbb{Z} / (12)$, | |
we see that $1 \otimes 1$ spans $\mathbb{Z} / (10) \otimes \mathbb{Z} / | |
(12)$ and this tensor | |
product is a cyclic group. | |
Note that | |
$1 \otimes 0 = 1 \otimes (10 \cdot 0) = 10 \otimes 0 = 0 \otimes 0 = 0$ | |
and | |
$0 \otimes 1 = (12 \cdot 0) \otimes 1 = 0 \otimes 12 = 0 \otimes 0 = 0$. | |
Now, | |
$10 (1 \otimes 1) = 10 \otimes 1 = 0 \otimes 1 = 0$ | |
and | |
$12 (1 \otimes 1) = 1 \otimes 12 = 1 \otimes 0 = 0$, | |
so the cyclic group $\mathbb{Z} / (10) \otimes \mathbb{Z} / (12)$ has order | |
dividing both | |
$10$ and $12$. This means that the cyclic group has order dividing | |
$\gcd(10, 12) = 2$. | |
To show that the order of $\mathbb{Z} / (10) \otimes \mathbb{Z} / (12)$, | |
define a bilinear map | |
$g: \mathbb{Z} / (10) \times \mathbb{Z} / (12) \to \mathbb{Z} / (2)$ via | |
$g : (x, y) \mapsto xy$. The universal property of tensor products then | |
says that there is a unique linear map | |
$f: \mathbb{Z} / (10) \otimes \mathbb{Z} / (12) \to \mathbb{Z} / (2)$ making | |
the diagram | |
\[ | |
\xymatrix{ | |
\mathbb{Z} / (10) \times \mathbb{Z} / (12) \ar[r]^\otimes \ar[rd]_g | |
& \mathbb{Z} / (10) \otimes \mathbb{Z} / (12) \ar[d]^f \\ | |
& \mathbb{Z} / (2). | |
} | |
\] | |
commute. In particular, this means that $f (x \otimes y) = g(x, y) = xy$. | |
Hence, $f(1 \otimes 1) = 1$, so $f$ is surjective, and therefore, | |
$\mathbb{Z} / (10) \otimes \mathbb{Z} / (12)$ has size at least two. This | |
allows us to | |
conclude that $\mathbb{Z} / (10) \otimes \mathbb{Z} / (12) = \mathbb{Z} / (2)$. | |
\end{example} | |
We now generalize the above example to tensor products of cyclic groups. | |
\begin{example} | |
Let $d=\gcd(m,n)$. We will show that | |
$(\mathbb{Z}/m\mathbb{Z})\otimes(\mathbb{Z}/n\mathbb{Z})\simeq(\mathbb{Z}/d\mathbb{Z})$, | |
and thus in particular if $m$ and $n$ | |
are relatively prime, then | |
$(\mathbb{Z}/m\mathbb{Z})\otimes(\mathbb{Z}/n\mathbb{Z})\simeq(0)$. First, | |
note that | |
any $a\otimes b\in(\mathbb{Z}/m\mathbb{Z})\otimes(\mathbb{Z}/n\mathbb{Z})$ | |
can be written as $ab(1\otimes 1)$, | |
so that $(\mathbb{Z}/m\mathbb{Z})\otimes(\mathbb{Z}/n\mathbb{Z})$ is generated | |
by $1\otimes 1$ and hence | |
is a cyclic group. We know from elementary number theory that $d=xm+yn$ | |
for some $x,y\in\mathbb{Z}$. We have $m(1\otimes 1)=m\otimes 1=0\otimes | |
1=0$ and | |
$n(1\otimes 1)=1\otimes n=1\otimes0=0$. Thus $d(1\otimes 1)=(xm+yn)(1\otimes | |
1)=0$, so that $1\otimes1$ has order dividing $d$. | |
Conversely, consider the map | |
$f:(\mathbb{Z}/m\mathbb{Z})\times(\mathbb{Z}/n\mathbb{Z})\rightarrow(\mathbb{Z}/d\mathbb{Z})$ | |
defined by | |
$f(a+m\mathbb{Z},b+n\mathbb{Z})=ab+d\mathbb{Z}$. This is well-defined, | |
since if $a'+m\mathbb{Z}=a+m\mathbb{Z}$ | |
and $b'+n\mathbb{Z}=b+n\mathbb{Z}$ then $a'=a+mr$ and $b'=b+ns$ for some | |
$r,s$ and | |
thus $a'b'+d\mathbb{Z}=ab+(mrb+nsa+mnrs)+d\mathbb{Z}=ab+d\mathbb{Z}$ | |
(since $d=\gcd(m,n)$ | |
divides $m$ and $n$). This is obviously bilinear, and hence induces a map | |
$\tilde{f}:(\mathbb{Z}/m\mathbb{Z})\otimes(\mathbb{Z}/n\mathbb{Z})\rightarrow(\mathbb{Z}/d\mathbb{Z})$, | |
which | |
has $\tilde{f}(1\otimes1)=1+d\mathbb{Z}$. But the order of $1+d\mathbb{Z}$ | |
in $\mathbb{Z}/d\mathbb{Z}$ is $d$, so that the order of $1\otimes1$ in | |
$(\mathbb{Z}/m\mathbb{Z})\otimes(\mathbb{Z}/n\mathbb{Z})$ must be at least | |
$d$. Thus $1\otimes1$ is in fact | |
of order $d$, and the map $\tilde{f}$ is an isomorphism between cyclic groups | |
of order $d$. | |
\end{example} | |
Finally, we present an example involving the interaction of $\hom$ and the | |
tensor product. | |
\begin{example} | |
Given an $R$-module $M$, let us use the notation $M^* = \hom_R(M,R)$. | |
We shall define a functorial map | |
\[ M^* \otimes_R N \to \hom_R(M,N), \] | |
and show that it is an isomorphism when $M$ is finitely generated and free. | |
Define $\rho':M^*\times N\rightarrow\hom_R(M,N)$ by | |
$\rho'(f,n)(m)=f(m)n$ (note that $f(m)\in R$, and the multiplication $f(m)n$ | |
is that between an element of $R$ and an element of $N$). This is bilinear, | |
\[\rho'(af+bg,n)(m)=(af+bg)(m)n=(af(m)+bg(m))n=af(m)n+bg(m)n=a\rho'(f,n)(m)+b\rho'(g,n)(m)\] | |
\[\rho'(f,an_1+bn_2)(m)=f(m)(an_1+bn_2)=af(m)n_1+bf(m)n_2=a\rho'(f,n_1)(m)+b\rho'(f,n_2)(m)\] | |
so it induces a map $\rho:M^*\otimes N \rightarrow \hom(M,N)$ with | |
$\rho(f\otimes n)(m)=f(m)n$. This homomorphism is unique since the $f\otimes | |
n$ generate $M^*\otimes N$. \\ | |
\noindent Suppose $M$ is free on the set $\{a_1,\ldots,a_k\}$. Then | |
$M^*=\hom(M,R)$ is free on the set $\{f_i:M\rightarrow R,$ $ | |
f_i(r_1a_1+\cdots+r_ka_k)=r_i\}$, because there are clearly no | |
relations among the $f_i$ and because any $f:M\rightarrow R$ has | |
$f=f(a_1)f_1+\cdots+f(a_n)f_n$. Also note that any element $\sum h_j\otimes | |
p_j \in M^*\otimes N$ can be written in the form $\sum_{i=1}^k f_i\otimes | |
n_i$, by setting $n_i=\sum h_j(a_i)p_j$, and \textit{that this is unique} | |
because the $f_i$ are a basis for $M^*$.\\ | |
\noindent We claim that the map $\psi:\hom_R(M,N)\rightarrow M^*\otimes N$ | |
defined by $\psi(g)=\sum_{i=1}^k f_i\otimes g(a_i)$ is inverse to $\rho$. Given | |
any $\sum_{i=1}^k f_i\otimes n_i\in M^*\otimes N$, we have | |
\[\rho(\sum_{i=1}^k f_i\otimes n_i)(a_j)=\sum_{i=1}^k\rho(f_i\otimes | |
n_i)(a_j)=\sum_{i=1}^kf_i(a_j)n_i=n_j\] | |
Thus, $\rho(\sum_{i=1}^k f_i\otimes n_i)(a_i)=n_i$, and thus | |
$\psi(\rho(\sum_{i=1}^k f_i\otimes n_i))=\sum_{i=1}^k f_i\otimes n_i$. Thus, | |
$\psi\circ\rho=\id_{M^*\otimes N}$.\\ | |
\noindent Conversely, recall that for $g:M\rightarrow N\in\hom_R(M,N)$, | |
we defined $\psi(g)=\sum_{i=1}^k f_i\otimes g(a_i)$. Thus, | |
\[\rho(\psi(g))(a_j)=\rho(\sum_{i=1}^k f_i\otimes | |
g(a_i))(a_j)=\sum_{i=1}^k\rho(f_i\otimes g(a_i))(a_j)=\sum_{i=1}^k | |
f_i(a_j)g(a_i)=g(a_j)\] | |
and because $\rho(\psi(g))$ agrees with $g$ on the $a_i$, it is the | |
same element of $\hom_R(M,N)$ because the $a_i$ generate $M$. Thus, | |
$\rho\circ\psi=\id_{\hom_R(M,N)}$.\\ | |
\noindent Thus, $\rho$ is an isomorphism. | |
\end{example} | |
We now interpret localization as a tensor product. | |
\begin{proposition} \label{locisbasechange} | |
Let $R$ be a commutative ring, $S \subset R$ a multiplicative subset. Then | |
there | |
exists a canonical isomorphism of functors: | |
\[ \phi: S^{-1}M \simeq S^{-1 }R \otimes_R M . \] | |
\end{proposition} | |
\begin{proof} | |
Here is a construction of $\phi$. If $x/s \in S^{-1}M$ where $x \in M, s \in | |
S$, we define | |
\[ \phi(x/s) = (1/s) \otimes m. \] | |
Let us check that this is well-defined. Suppose $x/s = x'/s'$; then this means | |
there is $t \in S$ with | |
\[ xs't = x'st . \] | |
From this we need to check that $\phi(x/s) = \phi(x'/s')$, i.e. that $1/s | |
\otimes x$ and $1/s' \otimes x'$ represent the same elements in the tensor | |
product. But we know from the last statement that | |
\[ \frac{1}{ss't} \otimes xs't = \frac{1}{ss't} x'st \in S^{-1}R \otimes M \] | |
and the first is just | |
\[ s't( \frac{1}{ss't} \otimes x) = \frac{1}{s} \otimes x \] | |
by linearity, while the second is just | |
\[ \frac{1}{s'} \otimes x' \] | |
similarly. One next checks that $\phi$ is an $R$-module homomorphism, which we | |
leave to the reader. | |
Finally, we need to describe the inverse. The inverse $\psi: S^{-1}R \otimes M | |
\to S^{-1}M$ is easy to construct because it's a map out of the tensor product, | |
and we just need to give a bilinear map | |
\[ S^{-1} R \times M \to S^{-1}M , \] | |
and this sends $(r/s, m)$ to $mr/s$. | |
It is easy to see that $\phi, \psi$ are inverses to each other by the | |
definitions. | |
\end{proof} | |
It is, perhaps, worth making a small categorical comment, and offering an | |
alternative argument. | |
We are given two functors $F,G$ from $R$-modules to $S^{-1}R$-modules, where | |
$F(M) = S^{-1}R \otimes_R M$ and $G(M) = S^{-1}M$. | |
By the universal property, the map $M \to S^{-1}M$ from an $R$-module to a | |
tensor product gives a natural map | |
\[ S^{-1}R \otimes_R M \to S^{-1}M, \] | |
that is a natural transformation $F \to G$. | |
Since it is an isomorphism for free modules, it is an isomorphism for all | |
modules by a standard argument. | |
\subsection{Tensor products of algebras} | |
\label{tensprodalg} | |
There is one other basic property of tensor products to discuss before moving | |
on: namely, what happens when one tensors a ring with another ring. We shall | |
see that this gives rise to \emph{push-outs} in the category of rings, or | |
alternatively, coproducts in the category of $R$-algebras. | |
Let $R$ be a commutative ring and suppose $R_1, R_2$ are $R$-algebras. That is, we have ring homomorphisms | |
\( \phi_0: R \to R_0, \quad \phi_1: R \to R_1. \) | |
\begin{proposition} | |
$R_0 \otimes_R R_1$ has the structure of a commutative ring in a natural way. | |
\end{proposition} | |
Indeed, this | |
multiplication multiplies two typical elements $x \otimes y, x' \otimes y'$ of | |
the tensor product by | |
sending them to | |
$xx' \otimes yy'$. | |
The ring structure is determined by this formula. One ought to check that this | |
approach respects the relations of the tensor product. We will do so in an | |
indirect way. | |
\begin{proof} | |
Notice that giving a multiplication law on $R_0 \otimes_R R_1$ is equivalent to giving an $R$-bilinear map | |
\[ (R_0 \otimes_R R_1) \times (R_0 \otimes R_1) \to R_0 \otimes_R R_1,\] | |
i.e. an $R$-linear map | |
\[ (R_0 \otimes_R R_1) \otimes_R (R_0 \otimes R_1) \to R_0 \otimes_R R_1\] | |
which satisfies certain constraints (associativity, commutativity, etc.). | |
But the left side is isomorphic to $(R_0 \otimes_R R_0) \otimes_R (R_1 | |
\otimes_R R_1)$. Since we have bilinear maps $R_0 \times R_0 \to R_0$ and $R_1 | |
\times R_1 \to R_1$, we get linear maps | |
$R_0 \otimes_R R_0 \to R_0$ and $R_1 \otimes_R R_1 \to R_1$. | |
Tensoring these maps gives the multiplication as a bilinear map. It is easy to | |
see that these two approaches are the same. | |
We now need to check that this operation is commutative and associative, with | |
$1 \otimes 1$ as a unit; moreover, it distributes over addition. Distributivity | |
over addition is built into the construction (i.e. in view of bilinearity). The | |
rest (commutativity, associativity, units) can be checked directly on the | |
generators, since we have distributivity. | |
We shall leave the details to the reader. | |
\end{proof} | |
We can in fact describe the tensor product of $R$-algebras by a universal | |
property. We will | |
describe a commutative diagram: | |
\[ | |
\xymatrix{ | |
& R \ar[rd] \ar[ld] & \\ | |
R_0 \ar[rd] & & R_1 \ar[ld] \\ | |
& R_0 \otimes_R R_1 | |
} | |
\] | |
Here $R_0 \to R_0 \otimes_R R_1$ sends $x \mapsto x \otimes 1$; similarly for $R_1 | |
\mapsto R_0 \otimes_R R_1$. These are ring-homomorphisms, and it is easy to | |
see that | |
the above | |
diagram commutes, since $r \otimes 1 = 1 \otimes r = r(1 \otimes 1)$ for $r \in | |
R$. | |
In fact, | |
\begin{proposition} | |
$R_0 \otimes_R R_1$ is universal with respect to this property: in the language | |
of category theory, the above diagram is a pushout square. | |
\end{proposition} | |
This means for any commutative ring $B$, and every pair of maps $u_0: R_0 \to | |
B$ and $u_1: R_1 \to B$ such that the pull-backs $R \to R_0 \to B$ and $R \to | |
R_1 \to B$ are the same, then we get a unique map of rings | |
\[ R_0 \otimes_R R_1 \to B \] | |
which restricts on $R_0, R_1$ to the morphisms $u_0, u_1$ that we started with. | |
\begin{proof} If $B$ is a ring as in the previous paragraph, we make $B$ into an $R$-module by the map $R \to R_0 \to B$ (or | |
$R \to R_1 \to B$, it is the same by assumption). | |
This map $R_0 \otimes_R R_1 \to B$ sends | |
\[ x \otimes y \to u_0(x) u_1(y). \] | |
It is easy to check that $(x,y) \to u_0(x)u_1(y)$ is $R$-bilinear (because of | |
the condition that the two pull-backs of $u_0, u_1$ to $R$ are the same), and | |
that it gives a homomorphism of rings $R_0 \otimes_R R_1 \to B$ which | |
restricts to $u_0, u_1$ on $R_0, | |
R_1$. One can check, for instance, that this is a homomorphism of rings by | |
looking at the generators. | |
It is also clear that $R_0 \otimes_R R_1 \to B$ is unique, because we know | |
that the | |
map on elements of the form $x \otimes 1$ and $1 \otimes y$ is determined by | |
$u_0, u_1$; these generate $R_0 \otimes_R R_1$, though. | |
\end{proof} | |
In fact, we now claim that the category of rings has \emph{all} coproducts. We | |
see that the coproduct of any two elements exists (as the tensor product over | |
$\mathbb{Z}$). It turns out that arbitrary coproducts exist. More generally, | |
if $\left\{R_\alpha\right\}$ is a family of $R$-algebras, then one can define | |
an object | |
\[ \bigotimes_\alpha R_\alpha, \] | |
which is a coproduct of the $R_\alpha$ in the category of $R$-algebras. To do | |
this, we simply take the generators as before, as formal objects | |
\[ \bigotimes r_\alpha, \quad r_\alpha \in R_\alpha, \] | |
except that all but finitely many of the $r_\alpha$ are required to be the | |
identity. One quotients by the usual relations. | |
Alternatively, one may use the fact that filtered colimits exist, and | |
construct the infinite coproduct as a colimit of finite coproducts (which are | |
just ordinary tensor products). | |
\section{Exactness properties of the tensor product} | |
In general, the tensor product is not exact; it is only exact on the right, | |
but it can fail to preserve injections. Yet in some important cases it | |
\emph{is} | |
exact. We study that in the present section. | |
\subsection{Right-exactness of the tensor product} | |
We will start by talking about extent to which tensor products do preserve | |
exactness under any circumstance. | |
First, let's recall what is going on. If $M,N$ are $R$-modules over the | |
commutative ring $R$, we have defined another $R$-module $\hom_R(M,N)$ | |
of morphisms | |
$M \to N$. This is left-exact as a functor of $N$. In other words, if we fix | |
$M$ and let $N$ vary, then the construction of homming out of $M$ preserves | |
kernels. | |
In the language of category theory, this construction $N \to \hom_R(M,N)$ has | |
an adjoint. The other construction we discussed last time was this adjoint, | |
and it is the tensor | |
product. Namely, given $M,N$ we defined a \textbf{tensor product} $M \otimes_R | |
N$ such that giving a map $M \otimes_R N \to P$ into some $R$-module $P$ | |
is the same as giving a | |
bilinear map $\lambda: M \times N \to P$, which in turn is the same as giving | |
an $R$-linear map | |
\[ M \to \hom_R(N, P). \] | |
So we have a functorial isomorphism | |
\[ \hom_R(M \otimes_R N, P) \simeq \hom_R(M, \hom_R(N,P)). \] | |
Alternatively, tensoring is the left-adjoint to the | |
hom functor. By abstract nonsense, it follows that since $\hom(M, \cdot)$ | |
preserves cokernels, the left-adjoint preserves cokernels and is right-exact. | |
We shall see this directly. | |
\begin{proposition} | |
The functor $N \to M \otimes_R N$ is right-exact, i.e. preserves cokernels. | |
\end{proposition} | |
In fact, the tensor product is symmetric, so it's right exact in either | |
variable. | |
\begin{proof} | |
We have to show that if $N' \to N \to N'' \to 0$ is exact, then so is | |
\[ M \otimes_R N' \to M \otimes_R N \to M \otimes_R N'' \to 0. \] | |
There are a lot of different ways to think about this. For instance, we can | |
look at the direct construction. The tensor product is a certain quotient of a | |
free module. | |
$M \otimes_R N''$ is the quotient of the free module generated by $m \otimes | |
n'', m \in M, n \in N''$ modulo the usual relations. The map $M \otimes N \to | |
M \otimes N''$ sends $m \otimes n \to m \otimes n''$ if $n'' $ is the image of | |
$n$ in $N''$. Since each $n''$ can be lifted to some $n$, it is obvious that | |
the map $M \otimes_R N \to M \otimes_R N''$ is surjective. | |
Now we know that $M \otimes_R N''$ is a quotient of $M \otimes_R N$. But which | |
relations do you have to impose on $M \otimes_R N$ to get $M \otimes_R | |
N''$? In fact, each relation in $M \otimes_R N''$ | |
can be lifted to a relation in $M \otimes_R N$, but with some redundancy. So | |
the only thing to quotient | |
out by is the statement that $x \otimes y, x \otimes y'$ have the same image in | |
$M \otimes N''$. In particular, we have to quotient out by | |
\[ x \otimes y - x\otimes y' \ , y - y' \in N' \] | |
so that if we kill off $x \otimes n'$ for $n' \in N' \subset N$, then we get $M | |
\otimes N''$. This is a direct proof. | |
One can also give a conceptual proof. We would like to know that $M \otimes N''$ | |
is the cokernel of $M \otimes N' \to M \otimes N''$. In other words, we'd like | |
to know that if we mapped $M \otimes_R N$ into some $P$ and the pull-back to $M | |
\otimes_R N'$, it'd factor uniquely through $M \otimes_R N''$. | |
Namely, we need to show that | |
\[ \hom_R(M \otimes N'', P) = \ker(\hom_R(M \otimes N, P) \to \hom_R(M | |
\otimes N'', P)). \] | |
But the first is just $\hom_R(N'', \hom_R(M,P))$ by the adjointness property. | |
Similarly, the second is just | |
\[ \ker( \hom_R(N, \hom(M,P)) \to \hom_R(N', \hom_R(M,P)) \] | |
but this last statement is $\hom_R(N'', \hom_R(M,P))$ by just the statement | |
that $N'' = \mathrm{coker}(N ' \to N)$. | |
To give a map $N'' $ into some module (e.g. $\hom_R(M,P)$) is the same thing as | |
giving a map out of $N$ which kills $N'$. | |
So we get the functorial isomorphism. | |
\end{proof} | |
\begin{remark} | |
Formation of tensor products is, in general, \textbf{not} exact. | |
\end{remark} | |
\begin{example} \label{tensorbad} | |
Let $R = \mathbb{Z}$. Let $M = \mathbb{Z}/2\mathbb{Z}$. Consider the exact | |
sequence | |
\[ 0 \to \mathbb{Z} \to \mathbb{Q} \to \mathbb{Q}/\mathbb{Z} \to 0 \] | |
which we can tensor with $M$, yielding | |
\[ 0 \to \mathbb{Z}/2\mathbb{Z} \to \mathbb{Q} \otimes_{} | |
\mathbb{Z}/2\mathbb{Z} \to \mathbb{Q}/\mathbb{Z} \otimes | |
\mathbb{Z}/2\mathbb{Z} \to 0 \] | |
I claim that the second thing $\mathbb{Q} \otimes \mathbb{Z}/2\mathbb{Z}$ | |
is zero. This is because by tensoring with | |
$\mathbb{Z}/2\mathbb{Z}$, we've made multiplication by 2 identically zero. By | |
tensoring with $\mathbb{Q}$, we've made multiplication by 2 invertible. The | |
only way to reconcile this is to have the second term zero. In particular, the | |
sequence becomes | |
\[ 0 \to \mathbb{Z}/2\mathbb{Z} \to 0 \to 0 \to 0 \] | |
which is not exact. | |
\end{example} | |
\begin{exercise} | |
Let $R$ be a ring, $I, J \subset R$ ideals. Show that $R/I \otimes_R R/J | |
\simeq R/(I+J)$. | |
\end{exercise} | |
\subsection{A characterization of right-exact functors} | |
Let us consider additive functors on the category of $R$-modules. So far, | |
we know a very easy way of getting such functors: given an $R$-module $N$, we | |
have a functor | |
\[ T_N: M \to M \otimes_R N. \] | |
In other words, we have a way of generating a functor on the category of | |
$R$-modules for each $R$-module. These functors are all right-exact, as we | |
have seen. | |
Now we will prove a converse. | |
\begin{proposition} | |
Let $F$ be a right-exact functor on the category of $R$-modules that commutes | |
with direct sums. Then $F$ is isomorphic to some $T_N$. | |
\end{proposition} | |
\begin{proof} | |
The idea is that $N$ will be $F(R)$. | |
Without the right-exactness hypothesis, we shall construct a natural morphism | |
\[ F(R) \otimes M \to F(M) \] | |
as follows. Given $m \in M$, there is a natural map $R \to M$ sending $1 \to | |
m$. This identifies $M = \hom_R(R, M)$. But functoriality gives a map $F(R) | |
\times \hom_R(R, M) \to F(M)$, which is clearly $R$-linear; the universal | |
property of the tensor product now produces the desired transformation | |
$T_{F(R)} \to F$. | |
It is clear that $T_{F(R)}(M) \to F(M)$ is an isomorphism for $M = R$, and | |
thus for $M$ free, as both $T_{F(R)}$ and $F$ commute with direct sums. Now | |
let $M$ be any $R$-module. There is a ``free presentation,'' that is an exact | |
sequence | |
\[ R^I \to R^J \to M \to 0 \] | |
for some sets $I,J$; we get a commutative, exact diagram | |
\[ \xymatrix{ | |
T_{F(R)}(R^I)\ar[d] \ar[r] & T_{F(R)} (R^J) \ar[d] \ar[r] & T_{F(R)} (M) \ar[d] \ar[r] & 0 \\ | |
F(R^I) \ar[r] & F(R^J) \ar[r] & F( M )\ar[r] & 0 | |
}\] | |
where the leftmost two vertical arrows are isomorphisms. A diagram chase now | |
shows that $T_{F(R)}(M) \to F(M)$ is an isomorphism. In particular, $F \simeq | |
T_{F(R)}$ as functors. | |
\end{proof} | |
Without the hypothesis that $F$ commutes with arbitrary direct sums, we could only draw | |
the same conclusion on the category of \emph{finitely presented} modules; the | |
same proof as above goes through, though $I$ and $J$ are required to be | |
finite.\footnote{Recall that an additive functor commutes with finite direct | |
sums.} | |
\begin{proposition} | |
Let $F$ be a right-exact functor on the category of finitely presented $R$-modules that commutes | |
with direct sums. Then $F$ is isomorphic to some $T_N$. | |
\end{proposition} | |
From this we can easily see that localization at a multiplicative subset $S | |
\subset R$ is given by tensoring with $S^{-1}R$. Indeed, localization is a | |
right-exact functor on the category of $R$-modules, so it is given by | |
tensoring with some module $M$; applying to $R$ shows that $M=S^{-1}R$. | |
\subsection{Flatness} | |
In some cases, though, the tensor product is exact. | |
\begin{definition} \label{flatdefn} | |
Let $R$ be a commutative ring. An $R$-module $M$ is called \textbf{flat} if the | |
functor $N \to M \otimes_R N$ is exact. An $R$-algebra is \textbf{flat} if it is flat as an | |
$R$-module. | |
\end{definition} | |
We already know that tensoring with anything is right exact, | |
so the only thing to be checked for flatness of $M$ is that the operation of tensoring by $M$ | |
preserves injections. | |
\begin{example} | |
$\mathbb{Z}/2\mathbb{Z}$ is not flat as a $\mathbb{Z}$-module by | |
\rref{tensorbad}. | |
\end{example} | |
\begin{example} \label{projmoduleisflat} | |
If $R$ is a ring, then $R$ is flat as an $R$-module, because tensoring by $R$ | |
is the identity functor. | |
More generally, if $P$ is a projective module (i.e., homming out of $P$ | |
is exact), then $P$ is flat. | |
\end{example} | |
\begin{proof} | |
If $P = \bigoplus_A R$ is free, then tensoring with $P$ corresponds to taking | |
the direct sum $|A|$ times, i.e. | |
\[ P \otimes_R M = \bigoplus_A M. \] | |
This is because tensoring with $R$ preserves (finite or direct) infinite sums. | |
The functor $M \to \bigoplus_A M$ is exact, so free | |
modules are flat. | |
A projective module, as discussed earlier, is a direct summand of a free | |
module. So if $P$ is projective, $P \oplus P' \simeq \bigoplus_A R$ for some | |
$P'$. Then we have that | |
\[ (P \otimes_R M) \oplus (P' \otimes_R M) \simeq \bigoplus_A M. \] | |
If we had an injection $M \to M'$, then there is a direct sum decomposition | |
yields a diagram of maps | |
\[ \xymatrix{ | |
P \otimes_R M \ar[d] \ar[r] & \bigoplus_A M \ar[d] \\ | |
P \otimes_R M' \ar[r] & \bigoplus_A M' | |
}.\] | |
A diagram-chase now shows that the vertical map is injective. Namely, the | |
composition $P \otimes_R M \to \bigoplus_A M'$ is injective, so the vertical | |
map has to be injective too. | |
\end{proof} | |
\begin{example} | |
If $S \subset R$ is a multiplicative subset, then $S^{-1}R $ is a flat $R$-module, because localization is an | |
exact functor. | |
\end{example} | |
Let us make a few other comments. | |
\begin{remark} | |
Let $\phi: R \to R'$ be a homomorphism of rings. Then, first of all, any | |
$R'$-module can be regarded as an $R$-module by composition with $\phi$. In | |
particular, $R'$ is an $R$-module. | |
If $M$ is an $R$-module, we can define | |
\[ M \otimes_R R' \] | |
as an $R$-module. But in fact this tensor product is an $R'$-module; it has | |
an action of $R'$. If $x \in M$ and $a \in R'$ and $b \in R'$, multiplication | |
of $(x \otimes a) \in M \otimes_R R'$ by $b \in R'$ sends this, \emph{by | |
definition}, to | |
\[ b(x \otimes a) = x \otimes ab. \] | |
It is easy to check that this defines an action of $R'$ on $M \otimes_R R'$. | |
(One has to check that this action factors through the appropriate relations, | |
etc.) | |
\end{remark} | |
The following fact shows that the hom-sets behave nicely with respect to flat | |
base change. | |
\begin{proposition} | |
Let $M$ be a finitely presented $R$-module, $N$ an $R$-module. Let $S$ be a | |
flat $R$-algebra. Then the natural map | |
\[ \hom_R(M,N) \otimes_R S \to \hom_S( M \otimes_R S, N \otimes_R S) \] | |
is an isomorphism. | |
\end{proposition} | |
\begin{proof} | |
Indeed, it is clear that there is a natural map | |
\[ \hom_R(M, N) \to \hom_S(M \otimes_R S, N \otimes_R S) \] | |
of $R$-modules. The latter is an $S$-module, so the universal property gives | |
the map $\hom_R(M, N) \otimes_R S \to \hom_S(M \otimes_R S, N \otimes_R S)$ as | |
claimed. | |
If $N$ is fixed, then we have two contravariant functors | |
in $M$, | |
\[ T_1(M) = \hom_R(M, N) \otimes_R S, \quad T_2(M) = \hom_S(M \otimes_R S, N | |
\otimes_R S). \] | |
We also have a natural transformation $T_1(M) \to T_2(M)$. | |
It is clear that if $M$ is \emph{finitely generated} and \emph{free}, then the | |
natural transformation is an isomorphism (for example, if $M = R$, then we just | |
have the map $N \otimes_R S \to N \otimes_R S$). | |
Note moreover that both functors are left-exact: that is, given an exact | |
sequence | |
\[ M' \to M \to M'' \to 0, \] | |
there are induced exact sequences | |
\[ 0 \to T_1(M'') \to T_1(M) \to T_1(M'), \quad 0 \to T_2(M'') \to T_2(M) \to | |
T_2(M') .\] | |
Here we are using the fact that $\hom$ is always a left-exact functor and the | |
fact that tensoring with $S$ preserves exactness. (Thus it is here that we use | |
flatness.) | |
Now the following lemma will complete the proof: | |
\begin{lemma} | |
Let $T_1, T_2$ be contravariant, left-exact additive functors from the category of | |
$R$-modules to the category of abelian groups. Suppose a natural transformation | |
$t: T_1(M) \to T_2(M)$ is given, and suppose this is an isomorphism whenever | |
$M$ is finitely generated and free. Then it is an isomorphism for any finitely | |
presented module $M$. | |
\end{lemma} | |
\begin{proof} | |
This lemma is a diagram chase. Fix a finitely presented $M$, and choose a | |
presentation | |
\[ F' \to F \to M \to 0, \] | |
with $F', F$ finitely generated and free. | |
Then we have an exact and commutative diagram | |
\[ \xymatrix{ | |
0 \ar[r] & T_1(M) \ar[d]^{} \ar[r] & T_1(F) \ar[d]^{\simeq} \ar[r] & | |
T_1(F') \ar[d]^{\simeq} \\ | |
0 \ar[r] & T_2(M) \ar[r] & T_2(F) \ar[r] & T_2(F') . | |
}\] | |
By hypotheses, the two vertical arrows to the right are isomorphisms, as | |
indicated. A diagram chase now shows that the remaining arrow is an | |
isomorphism, which is what we wanted to prove. | |
\end{proof} | |
\end{proof} | |
\begin{example} | |
Let us now consider finitely generated flat modules over a principal ideal | |
domain $R$. By \rref{structurePID}, we know that any such $M$ is isomorphic to a | |
direct sum $\bigoplus R/a_i$ for some $a_i \in R$. But if any of the $a_i$ is | |
not zero, then that $a_i$ would be a nonzero zerodivisor on $M$. However, we | |
know no element of $R - \left\{0\right\}$ can be a zerodivisor on $M$. It | |
follows that all the $a_i = 0$. In particular, we have proved: | |
\begin{proposition} | |
A finitely generated module over a PID is flat if and only if it is free. | |
\end{proposition} | |
\end{example} | |
\subsection{Finitely presented flat modules} | |
In \cref{projmoduleisflat}, we saw that a projective module over any ring $R$ | |
was automatically flat. In general, the converse is flat. For instance, | |
$\mathbb{Q}$ is a flat $\mathbb{Z}$-module (as tensoring by $\mathbb{Q}$ is a | |
form of localization). However, because $\mathbb{Q}$ is divisible (namely, | |
multiplication by $n$ is surjective for any $n$), $\mathbb{Q}$ cannot be a free | |
abelian group. | |
Nonetheless: | |
\begin{theorem} \label{fpflatmeansprojective} | |
A finitely presented flat module over a ring $R$ is projective. | |
\end{theorem} | |
\begin{proof} | |
We follow \cite{We95}. | |
Let us define the following contravariant functor from $R$-modules to $R$-modules. | |
Given $M$, we send it to $M^* = \hom_\mathbb{Z}(M, \mathbb{Q}/\mathbb{Z})$. | |
This is made into an $R$-module in the following manner: given $\phi: M \to | |
\mathbb{Q}/\mathbb{Z}$ (which is just a homomorphism of abelian groups!) and $r | |
\in R$, we send this to $r\phi$ defined by $(r\phi)(m) = \phi(rm)$. | |
Since $\mathbb{Q}/\mathbb{Z}$ is an injective abelian group, we see that $M | |
\mapsto M^*$ is an \emph{exact} contravariant functor from $R$-modules to | |
$R$-modules. | |
In fact, we note that $0 \to A \to B \to C \to 0$ is exact implies $0 \to C^* \to B^* \to A^* \to 0$ is exact. | |
Let $F$ be any $R$-module. There is a natural homomorphism | |
\begin{equation} \label{twoduals} M^* \otimes_R F \to \hom_R(F, M)^*. | |
\end{equation} | |
This is defined as follows. Given $\phi: M \to \mathbb{Q}/\mathbb{Z}$ and $x \in | |
F$, we define a new map $\hom(F, M) \to \mathbb{Q}/\mathbb{Z}$ by sending a | |
homomorphism $\psi: F \to M$ to $\phi(\psi(x))$. | |
In other words, we have a natural map | |
\[ \hom_{\mathbb{Z}}(M, \mathbb{Q}/\mathbb{Z} ) \otimes_R F \to | |
\hom_{\mathbb{Z}}( \hom_R(F, M)^*, \mathbb{Q}/\mathbb{Z}). \] | |
Now fix $M$. | |
This map \eqref{twoduals} is an isomorphism if $F$ is \emph{finitely | |
generated} and free. | |
Both are right-exact (because dualizing is contravariant-exact!). | |
The ``finite presentation trick'' now shows that the map is an isomorphism if | |
$F$ is finitely presented. | |
\add{this should be elaborated on} | |
Fix now $F$ finitely presented and flat, and consider the above two quantities | |
in \eqref{twoduals} as functors in $M$. | |
Then the first functor is exact, so the second one is too. | |
In particular, $\hom_R(F, M)^*$ is an exact functor in $M$; in particular, if | |
$M \twoheadrightarrow M''$ is a surjection, then | |
\[ \hom_R(F, M'')^* \to \hom_R(F, M)^* \] | |
is an injection. But this implies that | |
\[ \hom_R(F, M) \to \hom_R(F, M'') \] | |
is a \emph{surjection,} i.e. that $F$ is projective. | |
Indeed: | |
\begin{lemma} $ A \to B \to C $ is exact if and only if $C^* \to B^* \to A^* $ is exact. | |
\end{lemma} | |
\begin{proof} | |
Indeed, one direction was already clear (from $\mathbb{Q}/\mathbb{Z}$ being an | |
injective abelian group). | |
Conversely, we note that $M = 0$ if and only if $M^* = 0$ (again by | |
injectivity and the fact that $(\mathbb{Z}/a)^* \neq 0$ for any $a$). | |
Thus dualizing reflects isomorphisms: if a map becomes an isomorphism after | |
dualized, then it was an isomorphism already. From here it is easy to deduce | |
the result (by applying the above fact to the kernel and image). | |
\end{proof} | |
\end{proof} | |