-
Notifications
You must be signed in to change notification settings - Fork 0
/
1_introduction.tex
323 lines (251 loc) · 11.9 KB
/
1_introduction.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
% !TEX root = 6490.tex
\section{Introduction}
The latest version of these notes can be found online at
\url{https://github.com/dkmiller/algebraic-groups/releases/download/refs%2Fheads%2Fmaster/algebraic-groups.pdf}.
Comments and corrections are appreciated.
\subsection{Disclaimer}
These notes originated in the course MATH 6490: Linear algebraic groups and
their Lie algebras, taught by David Zywina at Cornell University. However, the
notes have been substantially modified since then, and are not an exact
reflection of the content and style of the original lectures. In particular,
the notes often switch from a somewhat elementary approach to a more
sheaf-theoretic approach. Any errors are solely the fault of the author.
\subsection{Notational conventions}
We follow Bourbaki in writing $\dN,\dZ,\dQ$\ldots for the natural numbers,
integers, rationals, \ldots. The natural numbers are $\dN=\{1,2,\ldots\}$.
If $A$ is a commutative ring, $M$ is an $A$-module, and $a\in A$,
we write $M/a$ for $M/(a M)$. In particular, $A/a$ is the quotient of $A$ by
the ideal generated by $a$. All abelian groups will be treated as
$\dZ$-modules. So $A$ is an abelian group and $n\in \dZ$, the quotient
$A/n$ means $A/n\cdot A$ even if $A$ is written multiplicatively.
The notation $X_{/S}$ will mean ``$X$ is a scheme over $S$.'' If
$S=\spectrum(A)$, we will write $X_{/A}$ to mean that $X$ is a scheme over
$\spectrum(A)$.
We write $\transpose x$ for the transpose of a matrix $x$.
Examples are closed with a triangle $\triangleright$.
Content that can be skipped will be delimited by a star $\star$. Usually
this material will be much more advanced.
\subsection{Main references}
The standard texts are \cite{borel-1991,humphreys-1975,springer-2009}. In
these, the requisite algebraic geometry is done from scratch in an archaic
language. A good reference for modern (scheme-theoretic) algebraic geometry is
\cite{hartshorne-1977}, and a (very abstract) modern reference for algebraic
groups is the three volumes on group schemes \cite{sga3-i,sga3-ii,sga3-iii}
from the \emph{S\'eminaire de G\'eom\'etrie Alg\'ebrique}. A source lying
somewhat in the middle is Jantzen's book \cite{jantzen-2003}.
\subsection{A bestiary of examples}
Let $k$ be an algebraically closed field whose characteristic is \emph{not}
$2$. For example, $k$ could be $\dC$ or $\overline{\dF_p(t)}$. For now, we
define a \emph{linear algebraic group} over $k$ to be a subgroup
$G\subset \GL_n(k)$ defined by polynomial equations.
\begin{example}[General linear]
The archetypal example of an algebraic group is
$\GL_n(k)=\{g\in \matrices_n(k):g\text{ is invertible}\}$. As a subset of
$\GL_n(k)$, this is defined by the empty set of polynomial equations.
\end{example}
\begin{example}[Special linear]
Let $\SL_n(k)=\{g\in \GL_n(k):\det(g)=1\}$. This is defined by the equation
$\det(g)=1$.
\end{example}
\begin{example}[Orthogonal]\label{eg:1st-orthogonal}
Let $\operatorname{O}_n(k)=\{g\in \GL_n(k):g \transpose g=1\}$. This is cut
out by the equations
\[
\sum_{j=1}^n g_{i j} g_{k j} = \delta_{i k}
\]
for $1\leqslant i,k\leqslant n$.
\end{example}
\begin{example}[Special orthogonal]
This is $\SO_n(k)=\operatorname{O}_n(k)\cap \SL_n(k)$.
\end{example}
\begin{example}
This group doesn't have a special name, but for the moment we will write
$U_n(k)$ for the group of strictly upper triangular matrices:
\[
U_n(k) = \begin{pmatrix} 1 & \cdots & \ast \\ & \ddots & \vdots \\ & & 1 \end{pmatrix} \subset \GL_n(k) .
\]
Recall that a matrix $g\in \GL_n(k)$ is \emph{unipotent} if $(g-1)^m=0$ for
some $m\geqslant 1$. All elements of $U_n(k)$ are unipotent. The group $U_n(k)$
is defined by the equations
\[
\{g_{i j}=0\text{ for }j<i, g_{ii}=1\} .
\]
\end{example}
\begin{example}[Multiplicative]
We write $\Gm(k)=k^\times$ for the multiplicative group of $k$ with its obvious
group structure. Note that $\Gm=\GL(1)$.
\end{example}
\begin{example}[Additive]\label{eg:additive}
Write $\Ga(k)=k$, with its additive group structure. There is a natural
isomorphism $\varphi:\Ga\iso U_2$ given by $\varphi(x)=\begin{pmatrix} 1 & x \\ & 1 \end{pmatrix}$. Since
\[
\begin{pmatrix} 1 & x \\ & 1 \end{pmatrix} \begin{pmatrix} 1 & y \\ & 1 \end{pmatrix} = \begin{pmatrix} 1 & x+y \\ & 1 \end{pmatrix} ,
\]
this is a group homomorphism, and it is easy to see that $\varphi$ is an
isomorphism of the underlying varieties.
\end{example}
\begin{example}
For any $n\geqslant 0$, $\Ga^n(k)=k^n$ with the usual addition is a linear
algebraic group. We could embed it into $\GL_{2n}(k)$ via $2\times 2$ blocks
and the isomorphism $\Ga\iso U_2$ in \autoref{eg:additive}.
\end{example}
\begin{example}[Tori]
For any $n\geqslant 0$, we have a \emph{torus} of rank $n$, namely
\[
T(k) = \begin{pmatrix} \ast \\ & \ddots \\ & & \ast \end{pmatrix}\subset \GL_n(k) .
\]
This is clearly isomorphic to $\Gm^n(k)$. We call any linear algebraic group
isomorphic to some $\Gm^n$ a torus.
\end{example}
This list almost exhausts the class of simple algebraic groups over an
algebraically closed field. All of these groups make sense over
an arbitrary field. But over non-algebraically closed fields (or even base
rings that are not fields) thinking of algebraic groups in terms of their sets
of points doesn't work very well.
\subsection{Coordinate rings}
The standard references \cite{borel-1991,humphreys-1975,springer-2009} all treat
algebraic groups in terms of their sets of points in an algebraically closed
field. This leads to convoluted arguments, and (sometimes) theorems that are
actually wrong. It is better to use schemes. Since we will (almost) never use
non-affine schemes, we can study algebraic groups via their coordinate rings.
\begin{example}
Consider $G=\GL_n(k)$. For $g=(g_{i j})\in \matrices_n(k)$, we have
$g\in G$ if and only if $\det(g)\ne 0$. But this isn't an honest algebraic
equation. We can remedy this by noting that $\det(g)\ne 0$ if and only if
there exists $y\in k$ such that $\det(g)\cdot y=1$. Thus we can define the
\emph{coordinate ring} $k[G]$ of $G$, as
\[
k[G] = k[x_{i j},y] / (\det(x_{i j})y-1) .
\]
For $k$-algebras $A,B$, write $\hom_k(A,B)$ for the set of $k$-algebra
homomorphisms $A\to B$. There is a natural identification
\[
\hom_k(k[G],k) = \GL_n(k) .
\]
For $\varphi:k[G]\to k$, put $g_{i j} = \varphi(x_{i j})$ and
$b=\varphi(y)$. Then $\varphi$ is well-defined exactly if
$\det(g)\cdot b=1$. So $\varphi$ is uniquely determined by the choice of an
invertible matrix $g\in \GL_n(k)$. Since $b$ is determined by $g$ and $g$ can
be chosen arbitrarily in $\GL_n(k)$, this correspondence is a bijection. So in
some sense, $k[G]$ recovers $\GL_n(k)$.
\end{example}
Now let $k$ be an arbitrary field. For concreteness, you could think of one of
$\dQ$, $\dR$, $\dC$, or $\dF_p$. Put $A=k[x_{i j},y]/(\det(x_{i j}) y-1)$, and
define $\GL_n(R)=\hom_k(A,R)$ for any $k$-algebra $R$. We will think of
``$\GL(n)_{/k}$'' as $\spectrum(A)$, which is a topological space with structure
sheaf (essentially) $A$. The punchline is that the coordinate ring
$k[G]$ of an algebraic group $G$ determines everything we need to know about
$G$.
\begin{example}
Let $k$ be a field not of characteristic $2$. For $d\in k^\times$, let
$G_d\subset \dA_{/k}^2$ be the subscheme cut out by $x^2-d y^2=1$. In
other words, $k[G_d]=k[x,y]/(x^2-d y^2-1)$. The group operation is
\[
(x_1,y_1)\cdot (x_2,y_2) = (x_1 x_2+d y_1 y_2, x_1 y_2+x_2 y_1) .
\]
We would like to realize $G_d$ as a matrix group. Consider the map
$\varphi_d:G_d\to \GL(2)_{/k}$ given by
$(x,y)\mapsto \mat{x}{dy}{y}{x}$. As an exercise,
check that this is an isomorphism between $G_d$ and the subgroup
$\{g_{11}=g_{22},x_{12}=d x_{21}\}$ of $\GL(2)_{/k}$.
If $k=\bar k$, then consider the composite of $G_d\monic \GL(2)_{/k}\iso \GL(2)_{/k}$,
the second map being given by
\[
g\mapsto \begin{pmatrix} & \sqrt d \\ 1 \end{pmatrix} g \begin{pmatrix} & \sqrt d \\ 1 \end{pmatrix}^{-1} .
\]
It sends $(x,y)\in G_d$ to $\begin{pmatrix} a & b\sqrt d \\ b\sqrt d & a \end{pmatrix}$.
This has the same image as $\varphi_1:G_1\to \GL(2)_{/k}$. So if $k=\bar k$, then
$G_d\simeq G_1$.
\end{example}
\begin{hard}
Write $A_d=k[x]/(x^2-d)$. A more conceptual definition of $G_d$ is that it is
the restriction of scalars $G_d=\weil_{A_d/k} \Gm$. That is, for any
$k$-algebra $A$, we have $G_d(A) = \Gm(A\otimes_k A_d)$.
\end{hard}
\begin{example}
Set $k=\dR$. We claim that $G_1\not\simeq G_{-1}$. We give a topological
proof. The group $G_1(\dR)\subset \dA^2(\dR)$ is cut out by $x^2-y^2=1$, hence
non-compact. But $G_{-1}(\dR)=\{x^2+y^2=1\}$ is compact. Thus
$G_1\not\simeq G_{-1}$ over $\dR$.
But we have seen that $G_{-1}\simeq G_1$ ``over $\dC$.'' As an exercise,
convince yourself that $G_1\simeq \Gm$.
\end{example}
It turns out that ``twists of $\Gm$ over $k$ up to isomorphism'' are in
bijective correspondence with $k^\times/2$, via the correspondence
$d\mapsto G_d$.
\begin{hard}
This is easy to see. The $k$-forms of $\Gm$ is in natural bijection with
$\h^1(k,\automorphisms \Gm)=\h^1(k,\dZ/2)$. Kummer theory tells us that
$\h^1(k,\dZ/2)=k^\times/2$.
\end{hard}
\subsection{Structure theory for linear algebraic groups}\label{sec:str-thry-lag}
Let $G$ be linear algebraic group over $k$. There is a filtration of normal
subgroups
\begin{center}
\begin{tikzpicture}
\matrix (m) [
matrix of nodes,
nodes={anchor=west}
] {
& $G$ \\
& $\cup$ & finite\\
connected & $G^\circ$ \\
& $\cup$ & semisimple \\
solvable & $\rad G$ \\
& $\cup$ & torus \\
unipotent & $\urad G$ \\
& $\cup$ \\
& $1$ \\
};
\draw [decoration={brace,amplitude=0.5em},decorate]
(m-3-2.north -| m.east) -- (m-7-2.south -| m.east)
node [midway,xshift=1cm] {reductive};
\end{tikzpicture}
\end{center}
Here, $G^\circ$ is the \emph{identity component} (for the Zariski topology) of
$G$. The quotient $\pi_0(G)=G/G^\circ$ is a finite group. We write
$\rad G$ for the \emph{radical} of $G$, namely the maximal smooth connected
solvable normal subgroup of $G$. Finally, $\urad G$ is the \emph{unipotent
radical} of $G$, namely the largest smooth connected unipotent subgroup of
$G$. The quotient $G^\circ/\rad G$ is semisimple, and the quotient
$G^\circ/\urad G$ is reductive. In general, we say a connected algebraic
group $G$ is \emph{semisimple} if $\rad G=1$, and \emph{reductive} if
$\urad G=1$.
We've seen examples of tori and unipotent groups, and everybody knows plenty of
finite groups. Here is a more direct definition of semisimple groups that
allows us to give some basic examples.
\begin{example}[Semisimple]
Let $k=\dC$, and let $G$ be a connected linear
algebraic group over $k$. We say that $G$ is \emph{simple} if it is
non-commutative, and has no proper nontrivial closed normal subgroups. We say
$G$ is \emph{almost simple} if the only such subgroups are finite.
For example, $\SL(2)$ is almost simple, because its only nontrivial
closed normal subgroup is $\{\pm 1\}$.
A group $G$ is \emph{semisimple} if we have an isogeny
$G_1\times \cdots \times G_r\to G$ with the $G_i$ almost simple. Here an
\emph{isogeny} is a surjection with finite kernel.
\end{example}
Over $\dC$, the almost simple groups (up to isogeny) are:
\begin{center}
\begin{tabular}{c|c|c}
label & group & dimension \\ \hline
$\typeA_n$ $(n\geqslant 1)$ & $\SL(n+1)$ & $n^2-1$ \\
$\typeB_n$ $(n\geqslant 2)$ & $\SO(2n+1)$ & $n(2n+1)$\\
$\typeC_n$ $(n\geqslant 3)$ & $\Sp(2n)$ & $n(2n+1)$ \\
$\typeD_n$ $(n\geqslant 4)$ & $\SO(2n)$ & $n(2n-1)$
\end{tabular}
\end{center}
We make requirements on the index in these families to prevent degenerate
cases (e.g.~$\typeB_1=1$) or matching (e.g.~$\typeA_2=\typeC_2$). There are
five \emph{exceptional groups}
\begin{center}
\begin{tabular}{c|l}
label & dimension \\ \hline
$\typeE_6$ & $78$ \\
$\typeE_7$ & $133$ \\
$\typeE_8$ & $248$ \\
$\typeF_4$ & $52$ \\
$\typeG_2$ & $14$
\end{tabular}
\end{center}
Later on, we'll be able to understand why this list is complete.