-
Notifications
You must be signed in to change notification settings - Fork 4
/
02.tex
378 lines (334 loc) · 11.9 KB
/
02.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
\section{Lecture: 03/10/2024}
\begin{exercise}
If $A$ and $B$ are algebras, $M$ is a module over $A$ and $N$ is a module over $B$, then
$M\oplus N$ is a module over $A\times B$ with
\[
(a,b)\cdot (m,n)=(a\cdot m,b\cdot n).
\]
\end{exercise}
\index{Division algebra}
A \emph{division algebra} $D$ is an algebra such that every non-zero element
is invertible, that is for all $x\in D\setminus\{0\}$ there exists $y\in D$ such that $xy=yx=1$.
Modules over division algebras are very much like vector spaces. For example,
every finitely generated module $M$ over a division algebra has a basis.
Moreover, every linearly independent subset of
$M$ can be extended into a basis of $M$.
\begin{proposition}
Let $D$ be a division algebra, and $V$ be a finite-dimensional module over $D$. Then
$V$ is a simple module over $\End_D(V)$ and there exits $n\in\Z_{>0}$ such that
$\End_D(V)\simeq nV$ is semisimple.
\end{proposition}
\begin{proof}[Sketch of the proof]
Let $\{v_1,\dots,v_n\}$ be a basis of $V$. A direct calculation shows that the map
\[
\End_D(V)\to\bigoplus_{i=1}^nV=nV,\quad
f\mapsto (f(v_1),\dots,f(v_n)),
\]
is an injective homomorphism of $\End_D(V)$-modules.
Since
\[
\dim_D\End_D(V)=n^2=\dim_D(nV),
\]
it follows that the map is an isomorphism.
Thus
\[
\End_D(V)\simeq \bigoplus_{i=1}^nV.
\]
It remains to show that $V$ is simple. It is enough to prove that $V=\End_D(V)\cdot v=(v)$
for all $v\in V\setminus\{0\}$. Let $v\in V\setminus\{0\}$. If $w\in V$, then
there exists $f\in\End_D(V)$ such that $f\cdot v=f(v)=w$.
Thus $w\in (v)$ and therefore $V=(v)$.
\end{proof}
The proposition states that if $D$ is a division algebra, then
$D^{n}$ is a simple $M_n(D)$-module and that $M_n(D)\simeq n D^n$ as $M_n(D)$-modules.
\begin{exercise}
Let $M$, $N$, and $X$ be modules. Prove that
\begin{align}
\Hom_A(M\oplus N,X)\simeq\Hom_A(M,X)\times\Hom_A(N,X).
\end{align}
\end{exercise}
\begin{theorem}
Let $A$ be a finite-dimensional algebra and let
$S_1,\dots,S_k$ be the simple modules over $A$.
If
\[
M\simeq n_1S_1\oplus\cdots\oplus n_kS_k,
\]
then each $n_j$ is uniquely determined.
\end{theorem}
\begin{proof}
Since each $S_j$ is a simple module
and $S_i\not\simeq S_j$ if $i\ne j$,
Schur's lemma implies that
$\Hom_A(S_i,S_j)=\{0\}$ whenever $i\ne j$.
For each $j\in\{1,\dots,k\}$, routine calculations show that
\begin{align*}
\Hom_A(M,S_j) &\simeq \Hom_A\left(\bigoplus_{i=1}^k n_i S_i,S_j\right)
\simeq n_j\Hom_A(S_j,S_j).
\end{align*}
Since $M$ and $S_j$ are finite-dimensional vector spaces, it follows that
$\Hom_A(M,S_j)$ and $\Hom_A(S_j,S_j)$
are both finite-dimensional vector spaces.
Moreover, the identity $\id\colon S_j\to S_j$
is a module homomorphism and hence
% $\id\in\Hom_A(S_j,S_j)$ and hence
$\dim\Hom_A(S_j,S_j)\geq 1$.
Thus each $n_j$ is uniquely determined, as
\[
n_j=\frac{\dim\Hom_A(M,S_j)}{\dim\Hom_A(S_j,S_j)}.\qedhere
\]
\end{proof}
\begin{definition}
\index{Algebra!opposite}
If $A$ is an algebra, the \emph{opposite algebra} $A^{\op}$ is the vector space
$A$ with multiplication $A\times A\to A$, $(a,b)\mapsto ba=a\cdot_{\op}b$.
\end{definition}
An algebra $A$ is commutative if and only if $A=A^{\op}$.
\begin{lemma}
\label{lem:A^op}
If $A$ is an algebra, then $A^{\op}\simeq\End_A(A)$ as algebras.
\end{lemma}
\begin{proof}
Note that $\End_A(A)=\{\rho_a:a\in A\}$, where $\rho_a\colon
A\to A$, $x\mapsto xa$. Indeed, if $f\in\End_A(A)$, then
$f(1)=a\in A$. Moreover, $f(b)=f(b1)=bf(1)=ba$ and hence
$f=\rho_a$. The map
\[
A^{\op}\to \End_A(A),\quad a\mapsto\rho_a,
\]
is bijective and it is an algebra homomorphism, as
\[
\rho_a\rho_b(x)=\rho_a(\rho_b(x))=\rho_a(xb)=x(ba)=\rho_{ba}(x).\qedhere
\]
\end{proof}
\begin{lemma}
\label{lem:Mn_op}
If $A$ is an algebra and $n\in\Z_{>0}$, then $M_n(A)^{\op}\simeq
M_n(A^{\op})$ as algebras.
\end{lemma}
\begin{proof}
Let $\psi\colon M_n(A)^{\op}\to M_n(A^{\op})$, $X\mapsto X^T$,
where $X^T$ is the transpose matrix of $X$. Since $\psi$ is a bijective linear map, it is enough
to see that $\psi$ is a homomorphism. If $i,j\in\{1,\dots,n\}$, $a=(a_{ij})$ and $b=(b_{ij})$, then
\begin{align*}
(\psi(a)\psi(b))_{ij}&=\sum_{k=1}^n \psi(a)_{ik}\psi(b)_{kj}=\sum_{k=1}^n a_{ki}\cdot_{\op}b_{jk}\\
&=\sum_{k=1}^n b_{jk}a_{ki}=(ba)_{ji}=((ba)^T)_{ij}=\psi(a\cdot_{\op}b)_{ij}.\qedhere
\end{align*}
\end{proof}
\begin{lemma}
\label{lem:simple}
If $S$ is a simple module and $n\in\Z_{>0}$, then
\[
\End_A(nS)\simeq M_n(\End_A(S))
\]
as algebras.
\end{lemma}
\begin{proof}[Sketch of the proof]
Let $(\varphi_{ij})$ be a matrix with entries in $\End_A(S)$. We define a map
$nS\to nS$ as follows:
\[
\begin{pmatrix}
x_1\\
\vdots\\
x_n
\end{pmatrix}
\mapsto
\begin{pmatrix}
\varphi_{11} & \cdots & \varphi_{1n}\\
\cdots & \ddots & \vdots\\
\varphi{n1} & \cdots & \varphi_{nn}
\end{pmatrix}
\begin{pmatrix}
x_1\\
\vdots\\
x_n
\end{pmatrix}
=\begin{pmatrix}
\varphi_{11}(x_1)+\cdots+\varphi_{1n}(x_n)\\
\vdots\\
\varphi_{n1}(x_1)+\cdots+\varphi_{nn}(x_n)
\end{pmatrix}.
\]
The reader should prove that the map
\[
M_n(\End_A(S))\to\End_A(nS)
\]
is an injective algebra homomorphism.
It is surjective. Indeed, if $\psi\in\End_A(nS)$ and
$i,j\in\{1,\dots,n\}$ one defines $\psi_{ij}$ by
\[
\psi\begin{pmatrix}
x\\
0\\
\vdots\\
0
\end{pmatrix}
=\begin{pmatrix}
\psi_{11}(x)\\
\psi_{21}(x)\\
\vdots\\
\psi_{n1}(x)
\end{pmatrix},\dots,
\psi\begin{pmatrix}
0\\
0\\
\vdots\\
x
\end{pmatrix}
=\begin{pmatrix}
\psi_{1n}(x)\\
\psi_{2n}(x)\\
\vdots\\
\psi_{nn}(x)
\end{pmatrix}.\qedhere
\]
\end{proof}
\begin{exercise}
Prove Lemma \ref{lem:simple}.
\end{exercise}
\begin{exercise}
Let $M$, $N$, and $X$ be modules. Prove that
\begin{align}
\Hom_A(X,M\oplus N)\simeq\Hom_A(X,M)\times\Hom_A(X,N).
\end{align}
\end{exercise}
\begin{theorem}[Artin--Wedderburn]
\index{Artin--Wedderburn theorem}
Let $A$ be a finite-dimensional semisimple algebra with
$k$ isomorphism classes of simple modules. Then
\[
A\simeq M_{n_1}(D_1)\times\cdots\times M_{n_k}(D_k)
\]
for some $n_1,\dots,n_k\in\Z_{>0}$ and some division algebras $D_1,\dots,D_k$.
\end{theorem}
\begin{proof}
Decompose the regular representation as a sum of simple modules and
gather the simples by isomorphism classes to get
\[
A=\bigoplus_{i=1}^k n_iS_i,
\]
where each $S_i$ is simple and $S_i\not\simeq S_j$ whenever
$i\ne j$. Schur's lemma implies that
\begin{align*}
\End_A(A)\simeq\End_A\left(\bigoplus_{i=1}^kn_iS_i\right)
\simeq \prod_{i=1}^k\End_A(n_iS_i)
\simeq\prod_{i=1}^kM_{n_i}(\End_A(S_i)),
\end{align*}
where each $D_i=\End_A(S_i)$ is a division algebra by Schur's lemma.
Thus
\[
\End_A(A)\simeq\prod_{i=1}^kM_{n_i}(D_i).
\]
Since $\End_A(A)\simeq
A^{\op}$, it follows that
\begin{align*}
A=(A^{\op})^{\op}\simeq \prod_{i=1}^kM_{n_i}(D_i)^{\op}\simeq \prod_{i=1}^kM_{n_i}(D_i^{\op}).
\end{align*}
Since each $D_i$ is a division algebra, each $D_i^{\op}$ is also a division algebra.
\end{proof}
\begin{corollary}[Mollien]
\index{Mollien's theorem}
If $A$ is a finite-dimensional complex semisimple algebra
with $k$ isomorphism classes of simple modules,
then
\[
A\simeq\prod_{i=1}^k M_{n_i}(\C)
\]
for some $n_1,\dots,n_k\in\Z_{>0}$.
\end{corollary}
\begin{proof}
By Wedderburn's theorem,
\[
A\simeq \prod_{i=1}^k M_{n_i}(\End_A(S_i)^{\op}),
\]
where $S_1,\dots,S_k$ are representatives of the isomorphism classes of simple modules
and each $\End_A(S_i)$ is a division algebra. We claim that
\[
\End_A(S_i)=\{\lambda\id:\lambda\in\C\}\simeq\C
\]
for all $i\in\{1,\dots,k\}$. If
$f\in\End_A(S_i)$, then $f$ has an eigenvalue $\lambda\in\C$. Since
$f-\lambda\id$ is not an isomorphism, Schur's lemma implies that $f-\lambda\id=0$,
that is $f=\lambda\id$. Thus $\End_A(S_i)\to\C$, $f\mapsto\lambda$,
is an algebra isomorphism. In particular,
\[
A\simeq \prod_{i=1}^k M_{n_i}(\C).\qedhere
\]
\end{proof}
We conclude this section with a nice
application of semisimplicity.
\begin{exercise}
\label{xca:matrices}
Prove that there exists a
homomorphism $M_n(\R)\to M_m(\R)$ of $\R$-algebras
if and only
if $n$ divides $m$.
\end{exercise}
% \begin{exercise}
% Let $A$ and $B$ be algebras. Prove that the ideals of $A\times B$ are of the form
% $I\times J$, where $I$ is an ideal of $A$ and $J$ is an ideal of $B$.
% \end{exercise}
\subsection{Simple algebras and Wedderburn's theorem}
\begin{definition}
\index{Algebra!simple}
An algebra $A$ is said to be \emph{simple} if $A\ne\{0\}$ and $\{0\}$ and $A$ are the only ideals of $A$.
\end{definition}
\begin{proposition}
Let $A$ be a finite-dimensional simple algebra. There exists a non-zero left ideal
$I$ of minimal dimension. This ideal is a simple
$A$-module, and every simple $A$-module is isomorphic to $I$.
\end{proposition}
\begin{proof}
Since $A$ is finite-dimensional and $A$ is a left ideal of $A$, there exists a non-zero left ideal of minimal dimension. The minimality
of $\dim I$ implies that $I$ is a simple $A$-module.
Let $M$ be a simple $A$-module. In particular, $M\ne\{0\}$.
Since
\[
\Ann_A(M)=\{a\in A:a\cdot M=\{0\}\}
\]
is an ideal of $A$ and $1\in A\setminus\Ann_A(M)$, the simplicity of $A$ implies that
$\Ann_A(M)=\{0\}$ and hence $I\cdot M\ne\{0\}$ (because $I\cdot m= 0$ for all $m\in M$ yields
$I\subseteq\Ann_A(M)$ and $I$ is non-zero, a contradiction). Let $m\in M$ be such
that $I\cdot m\ne\{0\}$. The map
\[
\varphi\colon I\to M,\quad
x\mapsto x\cdot m,
\]
is a module homomorphism. Since $I\cdot m\ne\{0\}$, the map $\varphi$ is non-zero.
Since both $I$ and $M$ are simple, Schur's lemma implies that $\varphi$ is an isomorphism.
\end{proof}
If $D$ is a division algebra, then $M_n(D)$ is a simple algebra. The previous proposition
implies that the algebra $M_n(D)$ has a unique isomorphism class of simple modules. Each simple module
is isomorphic to $D^n$.
\begin{proposition}
Let $A$ be a finite-dimensional algebra.
If $A$ is simple, then $A$ is semisimple.
\end{proposition}
\begin{proof}
Let $S$ be the sum of the simple submodules appearing in the regular representation of $A$.
We claim that $S$ is an ideal of $A$. We know that $S$ is a left ideal, as the submodules of the regular representation
are exactly the left ideals of $A$. To show that $Sa\subseteq S$ for all $a\in A$ we need to prove that
$Ta\subseteq S$ for all simple submodule $T$ of $A$ and $a\in A$.
If $T\subseteq A$ is a simple submodule and $a\in A$,
let $f\colon T\to Ta$, $t\mapsto ta$. Since $f$ is a surjective
module homomorphism and $T$ is simple, it follows that
either $\ker f=\{0\}$ or $\ker f=T$. If $\ker f=T$, then
$f(T)=Ta=\{0\}\subseteq S$. If $\ker f=\{0\}$, then $T\simeq f(T)=Ta$ and hence $Ta$ is simple. Hence $Ta\subseteq S$.
Since $S$ is an ideal of $A$ and $A$
is a simple algebra, it follows either $S=\{0\}$ or $S=A$. Since $S\ne\{0\}$, because
there exists a non-zero left ideal $I$ of $A$ such that $I\ne\{0\}$ is of minimal dimension,
it follows that $S=A$, that is, the regular representation of $A$ is semisimple (because it is a sum of simple submodules). Therefore
$A$ is semisimple.
\end{proof}
\begin{theorem}[Wedderburn]
\index{Wedderburn's theorem}
Let $A$ be a finite-dimensional algebra. If $A$ is simple, then
$A\simeq M_n(D)$ for some $n\in\Z_{>0}$ and some division algebra $D$.
\end{theorem}
\begin{proof}
Since $A$ is simple, it follows that $A$ is semisimple. Artin--Wedderburn theorem implies that $A\simeq\prod_{i=1}^k M_{n_i}(D_i)$
for some $n_1,\dots,n_k$ and some division algebras $D_1,\dots,D_k$. Moreover, $A$ has
$k$ isomorphism classes of simple modules. Since $A$ is simple,
$A$ has only one isomorphism class of simple modules. Thus $k=1$ and hence
$A\simeq M_n(D)$ for some $n\in\Z_{>0}$ and some division algebra~$D$.
\end{proof}