-
Notifications
You must be signed in to change notification settings - Fork 0
/
Copy path2_Lie-algebras.tex
850 lines (721 loc) · 34.3 KB
/
2_Lie-algebras.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
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
% !TEX root = 6490.tex
\section{Lie algebras}
A Lie algebra is a linear object whose representation theory is (in
principle) manageable. To any linear algebraic group $G$ we will associate a
Lie algebra $\fg=\lie(G)$, and study the representation theory of $G$ via that
of $\fg$.
\subsection{Definition and first properties}
Fix a field $k$. Eventually we'll avoid characteristic $2$, and some theorems
will only be valid in characteristic zero.
\begin{definition}
A \emph{Lie algebra} over $k$ is a $k$-vector space $\fg$ equipped with a map
$[\cdot,\cdot]:\fg\times \fg\to \fg$ such that the following hold:
\begin{itemize}
\item $[\cdot,\cdot]$ is bilinear (it factors through $\fg\otimes\fg$).
\item $[x,x]=0$ for all $x\in \fg$ ($[\cdot,\cdot]$ factors through
$\bigwedge^2 \fg$).
\item The Jacobi identity $[x,[y,z]]+[y,[z,x]]+[z,[x,y]]=0$ holds for all
$x,y,z\in \fg$.
\end{itemize}
\end{definition}
We could have defined a Lie algebra over $k$ as being a $k$-vector space
$\fg$ together with $[\cdot,\cdot]:\bigwedge^2 \fg\to \fg$ satisfying the
Jacobi identity. We call $[\cdot,\cdot]$ the \emph{Lie bracket}.
The Lie bracket satisfies a number of basic properties. For example,
$[x,y]=-[y,x]$ because
\begin{align*}
0 &= [x+y,x+y] && \text{(alternating)} \\
&= [x,x]+[x,y]+[y,x]+[y,y] && \text{(bilinear)} \\
&= [x,y] + [y,x] . && \text{(alternating)}
\end{align*}
Later on, we'll use an alternate form of the Jacobi identity:
\begin{equation*}\label{eq:jac-ident}\tag{$*$}
[x,[y,z]]-[y,[x,z]] = [[x,y],z] .
\end{equation*}
\subsection{The main examples}
\begin{example}
Let $V$ be any $k$-vector space. Then the zero map $\bigwedge^2 V\to V$
trivially satisfies the Jacobi identity. We call any Lie algebra whose
bracket is identically zero \emph{commutative} (or \emph{abelian}).
\end{example}
\begin{example}[General linear]
Again, let $V$ be a $k$-vector space. The Lie algebra $\gl(V)=\End_k(V)$ as a
$k$-vector space, with bracket $[X,Y]=X\circ Y-Y\circ X$. It is a good exercise
(which everyone should do at least once in their life) to check that this
actually is Lie algebra. When $V=k^n$, we write $\gl_n(k)$ instead of
$\gl(k^n)$.
\end{example}
Often, we will just write $\gl_n$ instead of $\gl_n(k)$. We will also do this
for the other ``named'' Lie algebras.
\begin{example}[Special linear]
Put $\Sl_n(k)=\{X\in \gl_n(k):\trace x=0\}$. Since $\trace [x,y]=0$, this is a
Lie subalgebra of $\gl_n$. In fact, $[\gl_n,\gl_n]\subset \Sl_n$.
\end{example}
\begin{example}[Special orthogonal]
Put $\so_n(k) = \{x\in \gl_n(k):\transpose x=-x\}$. This is a Lie subalgebra
of $\gl_n(k)$ because if $x,y\in \so_n(k)$, then
\begin{align*}
\transpose{[x,y]}
&= \transpose y \transpose x-\transpose x \transpose y \\
&= (-y)(-x) - (-x)(-y) \\
&= -[x,y] .
\end{align*}
\end{example}
\begin{example}[Symplectic]
Let $J_n=\mat{}{-1_n}{1_n}{}\in \gl_{2n}(k)$. Put
\[
\fsp_{2n}(k) = \{x\in \gl_{2n}(k):J x+\transpose x J=0\} .
\]
As an exercise, check that this is a Lie subalgebra of $\gl_{2n}$.
\end{example}
As an exercise, show that $\dR^3$ with the bracket $[u,v]=u\times v$ (cross
product) is a Lie algebra over $\dR$.
\begin{theorem}[Ado]
Any finite dimensional Lie algebra over $k$ is isomorphic to a Lie subalgebra
of some $\gl_n(k)$.
\end{theorem}
\begin{proof}
In characteristic zero, this is in \cite[I \S 7.3]{bourbaki-lie-alg-1-3}. The
positive-characteristic case is dealt with in \cite[VI \S3]{jacobson-1979}.
\end{proof}
Another good exercise is to realize $\dR^3$ with bracket $[u,v]=u\times v$ as
a subalgebra of some $\gl_n(\dR)$.
\subsection{Homomorphisms and the adjoint representation}
\begin{definition}
A \emph{homomorphism} of Lie algebras $\varphi:\fg\to \fh$ is a $k$-linear map
such that $\varphi([x,y])=[\varphi x,\varphi y]$.
\end{definition}
If $\varphi:\fg\to \fh$ is a homomorphism of Lie algebras, then
$\fa=\ker(\varphi)$ is a Lie subalgebra of $\fg$. This is easy:
\begin{align*}
\varphi[x,y]
&= [\varphi(x),\varphi(y)] \\
&= [0,0] \\
&= 0 .
\end{align*}
\begin{definition}
A Lie subalgebra $\fa$ of $\fg$ is an \emph{ideal} if
$[\fa,\fg]\subset \fa$, i.e.~$[a,x]\in \fa$ for all $a\in \fa$, $x\in \fg$.
\end{definition}
Equivalently, $\fa\subset \fg$ is an ideal if $[\fg,\fa]\subset \fa$. If
$\fa\subset \fg$ is an ideal, we define the \emph{quotient algebra} $\fg/\fa$
to be $\fg/\fa$ as a vector space, with bracket induced by that of $\fg$. So
\[
[x+\fa,y+\fa] = [x,y]+\fa .
\]
Let's check that this makes sense. If $a_1,a_2\in \fa$, then
\begin{align*}
[x+a_1,x+a_2]
&= [x,y] + [x,a_2]+[a_1,y]+[a_1,a_2] \\
&\equiv [x,y] \pmod \fa .
\end{align*}
If $\varphi:\fg\to \fh$ is a Lie homomorphism, then we get an induced
isomorphism $\fg/\ker(\varphi)\iso \image(\varphi)\subset \fh$.
\begin{example}
Give $k$ the trivial Lie bracket. Then $\trace:\gl_n(k) \to k$ is a a
homomorphism. Indeed,
\[
\trace [x,y]=\trace(x y)-\trace(y x)= 0 .
\]
We could have defined $\Sl_n(k)=\ker(\trace)$; this is an ideal in
$\gl_n$, and $\gl_n(k)/\Sl_n(k)\iso k$.
\end{example}
\begin{definition}
Let $\fg$ be a Lie algebra over $k$. The \emph{adjoint representation} of $\fg$
is the map $\adjoint:\fg\to \gl(\fg)$ defined by $\adjoint(x)(y) = [x,y]$ for
$x,y\in \fg$.
\end{definition}
It is easy to check that $\adjoint:\fg\to \gl(\fg)$ is $k$-linear, that is
$\adjoint(c x)=c\adjoint(x)$ and $\adjoint(x+y)=\adjoint(x)+\adjoint(y)$.
It is bit less obvious how the adjoint action behaves with respect to
the Lie bracket. We compute
\begin{align*}
\adjoint([x,y])(z)
&= [[x,y],z] \\
&=_\eqref{eq:jac-ident} [x,[y,z]]-[y,[x,z]] \\
&= (\adjoint(x) \adjoint(y))(z) - (\adjoint(y)\adjoint(x))(z) .
\end{align*}
Thus $\adjoint[x,y] = [\adjoint(x),\adjoint(y)]$ and we have shown that
$\adjoint:\fg\to \gl(\fg)$ is a homomorphism of Lie algebras. If
$n=\dim_k(\fg)<\infty$, then $\adjoint:\fg\to \gl(\fg)\simeq \gl_n(k)$ is
an interesting finite-dimensional representation of $\fg$. If
$\dim(\fg)>1$, it cannot be surjective, and there are easy ways for it to
fail to be injective.
\begin{definition}
Let $\fg$ be a Lie algebra over $k$. The \emph{center} of $\fg$ is
\[
\zentrum(\fg) = \{x\in \fg:[x,y]=0\text{ for all }y\in \fg\} .
\]
\end{definition}
Note that $\zentrum(\fg)=\ker(\adjoint)$. There is a natural injection
$\fg/\zentrum(\fg)\monic \gl(\fg)$.
\begin{definition}
A Lie algebra $\fg$ is \emph{simple} if it is non-commutative, and has no
ideals except $0$ and $\fg$.
\end{definition}
Just as with simple algebraic groups, there is a classification theorem for
simple Lie algebras. Even better, since Lie algebras are more ``rigid'' than
algebraic groups, we don't have to worry about isogeny.
\begin{theorem}
Let $k$ be an algebraically closed field of characteristic zero. Up to
isomorphism, every Lie algebra is a member of the following list:
\begin{center}
\begin{tabular}{c|c}
$\typeA_n$ ($n\geqslant 1$) & $\Sl_{n+1}$ \\
$\typeB_n$ ($n\geqslant 2$) & $\so_{2n+1}$ \\
$\typeC_n$ ($n\geqslant 3$) & $\fsp_{2n}$ \\
$\typeD_n$ ($n\geqslant 4$) & $\so_{2n}$ \\
exceptional & $\fe_6$, $\fe_7$, $\fe_8$, $\ff_4$, $\fg_2$
\end{tabular}
\end{center}
\end{theorem}
The classification theorem for algebraic groups is proved via the
classification theorem for Lie algebras, which ends up being a matter of
combinatorics.
\subsection{Tangent spaces}\label{sec:tangent-space}
For the sake of clarity, we give a scheme-theoretic definition of algebraic
groups:
\begin{definition}
Let $S$ be a scheme. An \emph{algebraic group} over $S$ is an affine group
scheme of finite type over $S$.
\end{definition}
For the moment, we will say that an algebraic group $G$ over $S$ is
\emph{linear} if $G$ is a subgroup scheme of $\GL(\sL)$ for some locally
free sheaf $\sL$ on $S$.
Fix a field $k$, and let $G\subset \GL(n)_{/k}$ be a linear algebraic group cut
out by polynomials $f_1,\dots,f_r$. For any $k$-algebra $R$, we define
\[
G(R) = \{g\in \GL_n(R):f_1(g)=\cdots = f_r(g)=0\} .
\]
This is a functor $\algebras k\to \sets$. Since $G$ is a subgroup scheme of
$\GL(n)_{/k}$, the set $G(R)$ inherits the group structure from $\GL_n(R)$.
So we will think of $G$ as a functor $G:\algebras k \to \groups$. By the Yoneda
Lemma, the variety $G$ is determined by its functor of points
$G:\algebras k\to \sets$.
We are especially interested in the $k$-algebra of \emph{dual numbers},
$k[\varepsilon]=k[x]/x^2$. Informally, $\varepsilon$ should be thought of as a
``infinitesimal quantity'' in the style of Newton and Leibniz. The scheme
$\spectrum(k[\varepsilon])$ should be thought of as a ``point together with
a direction.''
\begin{hard}
For the moment, let $k$ be an arbitrary base ring, $X_{/k}$ a scheme. We define
the \emph{$n$-Jet space} of $X$ to be the scheme $\jet_n X$ whose functor of
points is $(\jet_n X)(A) = X(A[t]/t^{n+1})$. By \cite{vojta-2007}, this
functor is representable, so $\jet_n X$ is actually a scheme. The first jet
space $\jet_1 X$ is called the \emph{tangent space} of $X$, and denoted
$\tangent X$. The maps $A[t]/t^{n+1}\epic A[t]/t^n$ induce projections
$\jet_{n+1} X\to \jet_n X$. In particular, the tangent space of $X$ comes with
a canonical projection $\pi:\tangent X\to X$.
\end{hard}
Consider $G(k[\varepsilon])$. As a set, this consists of invertible $n\times n$
matrices with entries in $k[\varepsilon]$ on which the $f_i$ vanish. The map
$\varepsilon\mapsto 0$ from $k[\varepsilon]\to k$ induces a group homomorphism
$\pi:G(k[\varepsilon])\to G(k)$. We will consider this map as the ``tangent
space'' of $G(k)$. For each $g\in G(k)$, the fiber
$\pi^{-1}(g)$ is the ``tangent space of $G$ at $g$.''
Note that
\[
\pi^{-1}(g) = \{g+\varepsilon v:v\in \matrices_n(k):f_i(g+\varepsilon v)=0\text{ for all }i\} .
\]
For each $s$, the polynomial $f_s$ has a Taylor series expansion
\[
f_s(x_{i j}) = f_s(g) + \sum_{i,j} \frac{\partial f_s}{\partial x_{i,j}}(g) (x_{i,j}-g_{i,j}) .
\]
Since $f_s(g)=0$, we get
$f_s(g+\varepsilon v) = \varepsilon \sum_{i,j} \frac{\partial f_s}{\partial x_{i,j}}(g) v_{i,j}$.
It follows that
\[
\pi^{-1}(g) = \left\{g+\varepsilon v:v\in \matrices_n(k):\sum_{i,j} \frac{\partial f_s}{\partial x_{i j}}(g) v_{i j}=0\text{ for all }1\leqslant s\leqslant r\right\} .
\]
We will be especially interested in $\fg=\pi^{-1}(1)$. We will turn this into a
Lie algebra using the group structure on $G$.
\begin{hard}
From the scheme-theoretic perspective, the identity element $1\in G(k)$ comes
from a section $e:\spectrum(k)\to G$ of the structure $G\to \spectrum(k)$. The
\emph{scheme-theoretic Lie algebra} of $G$ is the fiber product
\[
\fg = e^\ast 1 = \tangent(G)\times_G \spectrum(k) .
\]
By definition, $\fg(A)=\ker\left(G(A[\varepsilon])\to G(A)\right)$ for any
$k$-algebra $A$. Just as with algebraic groups, we will write $\fg_{/k}$ for
$\fg$ thought of as a group scheme
over $k$, and just $\fg$ for $\fg(k)$.
\end{hard}
\subsection{Functors of points}
Recall that schemes can be thought of as functors $\algebras k\to \sets$. For
example, affine $n$-space is the functor $\dA^n(A)=A^n$. In fact, this is an
algebraic group, the operation coming from addition on $A$. Moreover, $\dA^n$
is \emph{representable} in the sense that
\[
\dA^n(A) = \hom_k(k[x_1,\dots,x_n],A) ,
\]
via $(a_1,\dots,a_n)\mapsto (x_i\mapsto a_i)$. Similarly,
$\SL(n)_{/k}$ sends a $k$-algebra $A$ to $\SL_n(A)$. It is represented by the
ring $k[x_{i j}]/(\det(x)-1)$. In general, we take some polynomials
$f_1,\dots,f_r\in k[x_1,\dots,x_n]$, and consider the subscheme
$X=V(f_1,\dots,f_r)\subset \dA^n$ which represents the functor
\[
X(A) = \{a\in A^n:f_1(a)=\cdots = f_r(a)=0\} .
\]
This has coordinate ring $k[x_1,\dots,x_n]/(f_1,\dots,f_r)$. Note that
\emph{any} finitely generated $k$-algebras is of this form. In other words, all
affine schemes of finite type over $k$ are subschemes of some affine space.
Clasically, one gives $X(\bar k)$ the \emph{Zariski topology}. If
$f\in k[X]=\dA^1(X)$, we get (by definition) a function
$f:X(\bar k)\to \bar k$. We define open subsets of $X(\bar k)$ to be
those of the form $\{x\in X(\bar k):f(x)\ne 0\}$. More scheme-theoretically,
we say that an \emph{open subscheme} of $X$ is the complement of any
$V(f)$ for $f\in k[X]$.
\begin{hard}
We can think of the Zariski topology as giving a Grothendieck topology on the
category of affine schemes over $k$. It is subcanonical, i.e.~all representable
functors are sheaves. So if we formally put $\affine k = (\algebras k)^\circ$,
then the category of schemes over $k$ embeds into
$\sheaves_\zariski(\affine k)$. When working with algebraic groups, often it is
better with the \'etale or fppf topologies. In particular, we will think of
quotients as living in $\sheaves_\fppf(\affine k)$.
\end{hard}
\begin{definition}
Let $k$ be a feld. A \emph{variety} over $k$ is a reduced separated scheme of
finite type over $k$.
\end{definition}
In particular, an affine variety is determined by a reduced $k$-algebra. We
can directly associate a ``geometric object'' to a $k$-algebra $A$, namely its
\emph{spectrum} $\spectrum(A)=\{\fp\subset A\text{ prime ideal}\}$. Finally, we
note that if $X=\spectrum(A)$ is of finite type over $k$ and if $k=\bar k$, then
$X(k)=\hom_k(A,k)$ is the set of maximal ideas in $A$.
\subsection{Affine group schemes and Hopf algebras}
An affine group $G_{/k}$ will give us a functor $G:\algebras k\to \groups$. As
such, it should have morphisms
\begin{align*}
m :G\times G &\to G && \text{``multiplication''} \\
e:1=\spectrum(k) &\to G && \text{``identity''} \\
i:G &\to G && \text{``inverse''}
\end{align*}
These correspond to ring homomorphisms
\begin{align*}
\Delta:A &\to A\otimes_k A && \text{``comultiplication''} \\
\varepsilon : A &\to k && \text{``counit''} \\
\sigma:A &\to A && \text{``coinverse''}
\end{align*}
Obviously the maps on groups satisfy some axioms like associativity etc. These
can be phrased by the commutativity of diagrams like
\[\begin{tikzcd}
G\times G\times G \ar[r, "\id\times m"] \ar[d, "m\times \id"]
& G\times G \ar[d, "m"] \\
G\times G \ar[r, "m"]
& G .
\end{tikzcd}\]
Most sources do not give all these diagrams explicitly. One that does is
the book \cite{waterhouse-1979}.
It is a good exercise to show that we have the following maps for the additive
and multiplicative groups:
\begin{center}
\begin{tabular}{c|cccc}
group & ring & comultiplication & coident. & coinverse \\ \hline
$\Ga$ & $k[x]$ & $x\mapsto x\otimes 1+1\otimes x$ & $x\mapsto 0$ & $x\mapsto -x$ \\
$\Gm$ & $k[x^{\pm 1}]$ & $x\mapsto x\otimes x$ & $x\mapsto 1$ & $x\mapsto x^{-1}$
\end{tabular}
\end{center}
The analogue of associativity for coordinate rings and comultiplication is the
commutativity of the following diagram:
\[\begin{tikzcd}
A \ar[r, "\Delta"] \ar[d, "\Delta"]
& A\otimes_k A \ar[d, "\id\otimes \Delta"] \\
A\otimes_k A \ar[r, "\Delta\otimes \id"]
& A\otimes_k A\otimes_k A .
\end{tikzcd}\]
The tuple $(A,\Delta,\varepsilon,\sigma)$ is called a \emph{Hopf algebra}. In
principle, all theorems about algebraic groups can be rephrased as theorems
about Hopf algebras, but this is not very illuminating. The one advantage is
that it is possible to speak of Hopf algebras for which the ring $A$ is
\emph{not} commutative. These show up in algebraic number theory, combinatorics,
and physics.
\begin{theorem}\label{thm:groups-linear}
If $G$ is an affine group scheme of finite type over $k$, then it is a
linear algebraic group over $k$.
\end{theorem}
\begin{proof}
This is \cite[VI\textsubscript{B} 11.11]{sga3-i}. Essentially, one looks at the
action of $G$
on the (possibly infinite-dimensional) vector space $k[G]$. A general theorem
yields a faithful finite-dimensional representation $V$, and we get $G$ as a
subgroup-scheme of $\GL(V)$.
\end{proof}
\begin{hard}
If $S$ is a dedekind scheme (noetherian, one-dimensional, regular) and $G_{/S}$
is a affine flat separated group scheme of finite type over $S$, then the
obvious generalization of \autoref{thm:groups-linear} holds for $G$; it is a
closed subgroup of $\GL(\sL)$ for $\sL$ a locally free sheaf on $S$
\cite[VI\textsubscript{B} 13.5]{sga3-i}.
Interestingly, this fails for more general base schemes.
\end{hard}
\subsection{Examples of group schemes}
So all affine group varieties over $k$ embed into $\GL(n)_{/k}$ \emph{as an
algebraic group}. We can think of an algebraic group in several ways:
\begin{enumerate}
\item As a group-valued functor on $\algebras k$.
\item As a commutative Hopf algebra over $k$.
\item As a subgroup of $\GL_n(k)$ cut out by poynomials.
\end{enumerate}
\begin{example}
The group of \emph{$n$-th roots of unity} is $\dmu_n$. Its coordinate ring is
$k[x]/(x^n-1)$. So
\[
\dmu_n(A) = \{a\in A:a^n=1\} ,
\]
with the group operation coming from multiplication in $A$. Note that
$\dmu_n\subset \Gm$. If the base field $k$ has characteristic
$p\nmid n$, then $\dmu_n(\bar k)$ is a cyclic group of order $n$. However, if
$n=p$, then $x^p-1=(x-1)^p$. So the coordinate ring
$k[x]/((x-1)^p)$ is non-reduced. This is our first example of a group scheme
that is \emph{not} a group variety. One realization of this is that
$\dmu_p(\overline{\dF_p})=1$. These problems can only occur in characteristic
$p>0$. If $G$ is a group scheme over a characteristic zero field, then
$G$ is automatically reduced.
Let $\phi:\Gm\to \Gm$ be the homomorphism $x\mapsto x^p$. Note that
$\ker(\phi)=\dmu_p$. On $\overline{\dF_p}$-points, this map is an isomorphism,
but $\phi$ is \emph{not} an isomorphism of algebraic groups.
\end{example}
Let $G_{/k}$ be a linear algebraic group. Recall that we defined the \emph{Lie
algebra} $\fg=\lie(G)$ of $G$ to be the functor on $k$-algebras given by
\[
\fg(A) = \ker\left(G(A[\varepsilon]/\varepsilon^2) \to G(A)\right) .
\]
In \autoref{sec:tangent-space}, we saw that if $G\subset \GL(n)_{/k}$ is cut out by
$f_1,\dots,f_r$, then the $k$-valued points could be computed as
\[
\fg = \left\{X\in \matrices_n(k):\sum_{i,j} \frac{\partial f_\alpha}{\partial x_{i j}}(1_n)\cdot X_{i j}=0\right\} .
\]
\begin{example}[Special linear]
Consider $\SL(n)_{/k}\subset \GL(n)_{/k}$, cut out by $\det=1$. Its Lie algebra
$\Sl_n=\lie(\Sl_n)$ is
\[
\Sl_n = \left\{1+x\varepsilon:x\in \matrices_n(k)\text{ and }\det(1+x\varepsilon)=1\right\} .
\]
The expansion of determinant is
$\det(1+x\varepsilon)=1+\trace(x)\varepsilon + O(\varepsilon^2)$. Thus
\[
\Sl_n = \left\{x\in \matrices_n :\trace(x)=0\right\} .
\]
As a functor on $k$-algebras, $\Sl_n(A)=\{x\in \matrices_n(A):\trace x=0\}$.
\end{example}
\begin{example}[Orthogonal]
Recall that $\Or(n)\subset \GL(n)$ is cut out by $g \transpose g=1$. So
\begin{align*}
\lie(\Or_n)
&= \{1+x\varepsilon : (1+x\varepsilon) \transpose{(1+x\varepsilon)}=1\} \\
&= \{1+X\varepsilon:x+\transpose X=0\} \\
&\simeq \{X\in \matrices_n :\transpose x=-x\} \\
&= \so_n .
\end{align*}
If the characteristic of the base field is not $2$, then
$\dim(\so_n)=n(n-1)/2$.
\end{example}
\subsection{Lie algebra of an algebraic group}\label{sec:lie-alg-of-group}
Fix a field $k$. Let $G_{/k}$ be a linear algebraic group. For any $k$-algebra
$A$, write $A[\varepsilon]=k[t]/t^2$. Then $A\mapsto G(A[\varepsilon])$ is a
new group functor, and we can define the \emph{Lie algebra} of $G$ to be
\[
\lie(G)(A) = \ker\left(G(A[\varepsilon]) \to G(A)\right) .
\]
A priori, this is a group functor. Often, we will write $\fg=\lie(G)$ to mean
$\lie(G)(k)$. If $G\subset \GL(n)$ is cut out by $f_1,\dots,f_r$, then
\[
\fg=\left\{1+\varepsilon x:x\in \matrices_n(k)\text{ and }f_\alpha(1+\varepsilon x)=0\text{ for all }\alpha\right\} .
\]
The group operation is addition on $x$,
i.e.~$(1+\varepsilon x)(1+\varepsilon y)=1+\varepsilon(x+y)$. Thus
$\fg$ is a $k$-vector space.
\begin{hard}
This section will contain many ``high-level'' interludes on the functorial
definition of $\lie(G)$. They are all strongly influenced by the exposition in
\cite[II \S3--4]{sga3-i}. Let $\sO:\algebras k\to \algebras k$ be the functor
$A\mapsto A$. Note that $\sO$ is represented by $k[t]$. Thus $\dA^1_{/k}$ has
the structure not only of a group scheme, but of a $k$-ring scheme, or
$(k,k)$-biring in the sense of \cite{borger-wieland-2005}. We define an
\emph{$\sO$-module} to be a functor $V:\algebras k\to \sets$ such that for each
$A$, the set $V(A)$ is given an $\sO(A)=A$-module structure in a functorial
way.
If $S:\algebras k \to \sets$ is a functor, we will write
$\schemes S$ for the category of representable functors
$X:\algebras k\to \sets$ together with a morphism $X\to S$. A morphism
$(X\to S)\to (Y\to S)$ in $\schemes S$ is a morphism of functors
$X\to Y$ making the following diagram commute:
\[\begin{tikzcd}
X \ar[r] \ar[dr]
& Y \ar[d] \\
& S .
\end{tikzcd}\]
There is an easy generalization of $\sO$ to a ring object
$\sO\in\schemes S$. It sends a scheme $X$ to the ring $\h^0(X,\sO_X)$ of
regular functions on $X$.
For a scheme $X$, the first jet space
$\jet_1 X$ is naturally an $\sO$-module in $\schemes X$. We could be fancy and
say that for any ring $A$, $A[\varepsilon]$ is naturally an abelian group
object in $\algebras A/A$, or we could define the $\sO$-module structure on
$\jet_1 X\xrightarrow\pi X$ directly. For $\varphi:\spectrum(A)\to X$, we need
to give
\[
(\jet_1 X)_{/X}(A) = \{f\in (\jet_1 X)(A):\pi\circ f = \varphi\}
\]
the structure of an $A$-module. Given $a\in A$, define
$\sigma_a:A[\varepsilon]\to A[\varepsilon]$ by
$\varepsilon\mapsto a\varepsilon$. This is a ring homomorphism, so it induces
$\sigma_a:\jet_1 X(A)\to \jet_1 X(A)$. The restriction of this map to
$(\jet_1 X)_{/X}(A)$ is the desired action of $A$. See \cite[II 3.4.1]{sga3-i}
for a more careful proof that this gives $(\jet_1 X)_{/X}$ the structure of
an $\sO$-module in $\schemes X$.
The point of all this is that if $M$ is an $\sO$-module in
$\schemes Y$ and $f:X\to Y$ is a morphism of schemes, then
$f^\ast M$, defined by
\[
(f^\ast M)(\spectrum(A)\xrightarrow p X) = M(\spectrum(A)\xrightarrow p X\xrightarrow f Y)
\]
is naturally an $\sO$-module in $\schemes X$. So if $G$ is an algebraic group,
$\jet_1 X$ is naturally an $\sO$-module in $\schemes G$, and the morphism
$e:1\to G$ gives $\fg=e^\ast \jet_1 G$ the structure of an $\sO$-module in
$\schemes k$.
\end{hard}
\begin{example}[Symplectic]
Let $J=\begin{pmatrix} & -1_n \\ 1_n \end{pmatrix}$. Recall that
\[
\Sp_{2n}(A) = \left\{g\in \GL_{2n}(A) : \transpose g J g=J\right\} .
\]
Let's compute the Lie algebra of $\Sp_{2n}$. For $X\in \matrices_{2n}(A)$, we
have
\begin{align*}
\transpose{(1+\varepsilon x)} J (1+\varepsilon x)
&= (1+\varepsilon\transpose x) J (1+\varepsilon x) \\
&= J+\varepsilon(\transpose x J+J x) .
\end{align*}
So
$\fsp_{2n}(A) \simeq \left\{x\in \matrices_{2n}(A):\transpose x J+J x = 0\right\}$.
\end{example}
Algebraic groups can be highly non-abelian, and so far all we've done is give
$\fg=\lie(G)$ the structure of a vector space. We'd like to give $\fg$ the
structure of a Lie algebra.
Let $f:G\to H$ be a homomorphism of algebraic groups. Then $f:G(A)\to H(A)$ is
a homomorphism for all $k$-algebras $A$. A morphism $f:G\to H$ corresponds
by the Yoneda Lemma to a unique element $\varphi\in \hom_k(k[H],k[G])$. For all
$A$, the induced map $G(A)\to H(A)$ is defined by
$\psi\mapsto \psi\circ\varphi$ via the identifications
$G(A)=\hom(k[G],A)$ and $H(A)=\hom(k[H],A)$.
\begin{example}
Recall that $\Gm(A)=A^\times = \hom(k[x^{\pm 1}], A)$. All homomorphisms
$f:\Gm\to \Gm$ come from $x\mapsto x^n$ for some $n\in \dZ$. On functors of
points, this is $a\mapsto a^n$ for $a\in A^\times$.
\end{example}
As before, let $f:G\to H$ be a morphism of algebraic groups. Let
$\fg=\lie(G)$, $\fh=\lie(G)$. There is an induced map
$\lie(f):\fg\to \fh$. On functors of points, it is induced by the
inclusions $\fg\subset \jet_1 G$, $\fh\subset \jet_1 H$ and the fact that
$f:\jet_1 G\to \jet_1 H$ is a group homomorphism. That is, for each
$k$-algebra $A$, $f_\ast:\fg(A)\to \fh(A)$ is induced by the commutative
diagram
\[\begin{tikzcd}
G(A[\varepsilon]) \ar[r, "f"]
& H(A[\varepsilon]) \\
\fg(A) \ar[r, "f_\ast"] \ar[u, hook]
& \fh(A). \ar[u, hook]
\end{tikzcd}\]
In particular, a representation $\rho:G\to \GL(n)$ induces a map on
Lie algebras $\fg\to \gl_n$.
\begin{hard}
The natural way of stating the above is that $\lie$ is a functor from group
schemes over $k$ to $\sO$-modules in $\schemes k$. What's more, if $G$ is a
group scheme and $\fg=\lie(G)$, then by \cite[II 3.3]{sga3-i},
there is a natural isomorphism of $A$-modules
$\fg(k)\otimes_k A\iso \fg(A)$. So the functor $\fg$ is actually determined
by the vector space $\fg(k)$. This justifies our passing between
$\fg$ and $\fg(k)$.
\end{hard}
Let's construct the Lie bracket on the Lie algebra of an algebraic group.
Recall that for a Lie algebra $\fg$, we defined a map
$\adjoint:\fg\to \gl(\fg) = \End_k(\fg)$ by $\adjoint(x)(y)=[x,y]$. The
bracket on $\fg$ is clearly determined by the adjoint map.
For $\fg=\lie(G)$, we'll define the adjoint map directly, then use this to
give $\fg$ the structure of a Lie algebra. For any $g\in G(A)$, we have an
action $\adjoint(g):G_A\to G_A$ given by $x\mapsto g x g^{-1}$ for any
$B\in \algebras A$. So we have a homomorphism of group functors
$\adjoint:G\to \automorphisms(G)$. In particular, $G(k)$ acts on $G$ by
$k$-automorphisms.
\begin{hard}
The general philosophy, due to Grothendieck and worked out carefully in
SGA, is that ``everything is a functor.'' That is, we think of every object
as living in the category of functors $\algebras k\to \sets$. For example,
if $X$ and $Y$ are schemes over $k$, then $\hom(X,Y)$ is the functor whose
\[
\hom(X,Y)(A) = \hom_{\schemes A}(X_A, Y_A) .
\]
We recover the usual hom-set as $\hom(X,Y)(k)$. Similarly, we define
$\automorphisms(G)(A)=\automorphisms_{\groups_A}(G_A)$. These functors are not
usually representable, but they are sheaves for all the reasonable Grothendieck
topologies on $\schemes k$. See \cite[I \S 1--3, II \S 1]{sga3-i} for a careful
exposition along these lines.
\end{hard}
By functoriality, there is a natural homomorphism
$\automorphisms(G)\to \automorphisms(\fg)$. Composing this with the map
$\adjoint:G\to \automorphisms(G)$ gives a map
$\adjoint:G\to \automorphisms(\fg)$. Explicitly, the map
$\adjoint:G(A)\to \automorphisms(\fg)(A)=\automorphisms(\fg_A)$ sends
$g$ to the automorphism $x\mapsto g x g^{-1}$ for any
$x\in G(B[\varepsilon])$, $B\in \algebras A$. It is easy to see that this
action respects $\sO$-module structure on $\fg$, so we in fact have a
representation $\adjoint:G\to \GL(\fg)$. Since
$\fg(A)=\fg(k)\otimes_k A$, we can think of $\fg$ as an honest
$k$-vector space and define $\GL(\fg)$ as the functor
$A\mapsto \automorphisms_A(\fg\otimes_k A)$. In either case, we have a
morphism $\adjoint:G\to \GL(\fg)$ of linear algebraic groups.
Apply $\lie(-)$ to the adjoint representation $\adjoint:G\to \GL(\fg)$. We
get a map $\adjoint:\fg\to \lie(\GL(\fg))=\gl(\fg)$. We claim that
$\fg$ together with $[x,y]=\adjoint(x)(y)$ is a Lie algebra.
\begin{example}
Let $G=\GL(n)$. We wish to verify that our functorial definition of
$\adjoint:\gl_n\to \gl(\gl_n)$ agrees with the elementary definition
$\adjoint(x)(y)=[x,y]$. Start with the action of $\GL(n)$ on
$\gl_n$. For any $k$-algebra $A$, this is the action by conjugation of
$\GL_n(A)\subset \GL_n(A[\varepsilon])$ on
$\gl_n(A)=\ker(\GL_n(A[\varepsilon])\to \GL_n(A))$. This is because
\[
g\cdot (1+\varepsilon x)(g^{-1}) = 1+\varepsilon \adjoint(g)(x) .
\]
So the functorial and elementary definitions of $\adjoint:\GL(n)\to \GL(\gl_n)$
agree.
Now let's differentiate $\adjoint:\GL(n)\to \GL(\gl_n)$. For
$A\in \algebras k$, we have a commutative diagram:
\[\begin{tikzcd}
\gl_n(A) \ar[r, hook] \ar[d, "\adjoint"]
& \GL_n(A[\varepsilon]) \ar[r, twoheadrightarrow] \ar[d, "\adjoint"]
& \GL_n(A) \ar[d, "\adjoint"] \\
\gl(\gl_n)(A) \ar[r, hook]
& \GL(\gl_n\otimes A[\varepsilon]) \ar[r, twoheadrightarrow]
& \GL(\gl_n\otimes A)
\end{tikzcd}\]
All that remains is the simple computation
\begin{align*}
\adjoint(1+ \varepsilon x)(y)
&= (1+\varepsilon x)y(1-\varepsilon x) \\
&= y+[x,y]\varepsilon
\end{align*}
So as an endomorphism of $\gl_n\otimes A[\varepsilon]$,
$\adjoint(1+\varepsilon x)$ is of the form $1+\varepsilon [x,-]$. So when we
identify $1+\varepsilon x$ with $x$, we get that $\adjoint(x)=[x,-]$ as
desired. See \cite[II 4.8]{sga3-i} for a proof in much greater generality.
\end{example}
This example yields the general fact that $\lie(G)$ is a Lie algebra. Choose an
embedding $G\monic \GL(n)$. By construction, the Lie functor is
limit-preserving, so $\fg\monic \gl_n$. The bracket on $\fg$ (as a subspace of
$\gl_n$ is induced by the bracket on $\gl_n$. We have seen that the functorial
definition of the bracket on $\gl_n$ matches the elementary definition, so
$\gl_n$ is a Lie algebra and $\fg$ is a Lie subalgebras of $\gl_n$.
Does $\lie(G)$ determine $G$? The answer is an easy no! For example,
$\SO(n)$ and $\Or(n)$ both have Lie algebra $\so_n$. Also
$\SL(n)$ and $\PSL(n)$ both have Lie algebra $\Sl_n$. There is are the
easy example $\lie(\Ga)=\lie(\Gm)=\gl_1$. Finally,
$\lie(G)=\lie(G^\circ)$, so $\lie(G)$ does not depend on the group
$\pi_0(G)=G/G^\circ$ of connected components of $G$. However, given a
semisimple Lie algebra $\fg$, there is a unique split connected,
simply-connected semisimple algebraic group $G$ with $\lie(G)=\fg$.
\begin{example}
Let's go over the Lie bracket on $\GL(n)$ in a more elementary manner. We have
\[
\gl_n
= \ker\left(\GL_n(k[\varepsilon])\to\GL_n(k)\right)
= \{1+\varepsilon x:x\in \matrices_n(k)\} \iso \matrices_n(k) .
\]
We'll write elements of $\matrices_n(k[\varepsilon])$ as $x+y\varepsilon$;
these should be thought of as a ``point'' $x$ together with an ``infinitesimal
direction'' $y$. The adjoint map is the homomorphism
$\adjoint:\GL(n)\to \GL(\gl_n)$, defined on $A$-points as the map
$\GL_n(A)\to \GL(\gl_n(A))$ given by $\adjoint(g)(x) = g x g^{-1}$. So for
$A=k[\varepsilon]$, the group $\GL_n(k[\varepsilon])$ acts on
$\matrices_n(k[\varepsilon])$ by conjugation. Take any
$1+\varepsilon b\in \GL_n(k[\varepsilon])$, and any
$x+\varepsilon y\in \matrices_n(k[\varepsilon])$. We compute:
\begin{align*}
(1+\varepsilon b)(x+\varepsilon y)(1-\varepsilon b)^{-1}
&= (x+\varepsilon y+\varepsilon b x)(1-\varepsilon b) \\
&= x+\varepsilon y + \varepsilon(b x-x b) \\
&= x+\varepsilon y + \varepsilon [b,x] \\
&= (1+\varepsilon[b,-])(x+\varepsilon y) .
\end{align*}
Thus $\adjoint(1+\varepsilon b)=1+\varepsilon [b,-]$ as an element of
$\End_k(\gl_n) = \gl(\gl_n)$. Summing it all up, we have a commutative
diagram:
\[\begin{tikzcd}
\gl_n \ar[r, "\adjoint"] \ar[d, "\wr"]
& \gl(\gl_n) \ar[d, "\wr"] \\
\matrices_n \ar[r, "\adjoint"]
& \End(\matrices_n)
\end{tikzcd}\]
so that $\adjoint(x)(y) = [x,y]$.
\end{example}
As we have already seen, $\lie(G)$ does not determine $G$. Even worse, the
functor $\lie$ is neither full nor faithful. It's not faithful because, for
example, if $\Gamma$ is a finite group and $G$ a connected group, then
$\automorphisms(\Gamma)\hookrightarrow \automorphisms(\Gamma\times G)$, but
since $(G\times\Gamma)^\circ\subset G$, any automorphism of $\Gamma$ acts
trivially on $\lie(\Gamma\times G)$. To see that $\lie$ is not faithful,
note that $\hom(\gl_1,\gl_1)$ is a one-dimensional vector space, while
$\hom(\Gm,\Gm)=\dZ$.
\begin{theorem}\label{thm:subgroup-lie-alg}
Let $k$ be a field of characteristic zero, $G_{/k}$ a linear algebraic group.
Let $\fg=\lie(G)$. The map $H\mapsto \lie(H)$ from connected closed subgroups
of $G$ to Lie subalgebras of $\fg$ is injective and preserves inclusions.
\end{theorem}
\begin{proof}
If $H_1\subset H_2$, use the functoriality of forming Lie algebras to see that
$\lie(H_1)\subset \lie(H_2)$ as subalgebras of $\fg$. The nontrivial part is to
show that $\lie(H)$, as a subalgebra of $\fg$, determines $H$.
Let $H_1$, $H_2$ be be two closed connected subgroups of $G$ such that
$\fh=\lie(H_1)=\lie(H_2)$. Then $H_3=(H_1\cap H_2)^\circ$ is a closed connected
subgroup of $G$. Since fiber products are computed pointwise and the Lie
functor is exact, we have $\lie(H_3)=\fh$. By \autoref{thm:cartier}, the
$H_i$ are all smooth, so $\dim(H_i)=\dim \fh$. But the only way
$H_3\subset H_1$ can both be smooth and irreducible with the same dimension is
for $H_1=H_3$. Similarly $H_2=H_3$, so $H_1=H_2$.
\end{proof}
\begin{example}
The subgroup
$\Gm\simeq \left\{\begin{pmatrix} a \\ & a^{-1} \end{pmatrix}\right\}$
and $\Ga\simeq \begin{pmatrix} 1 & \ast \\ & 1 \end{pmatrix}$ of $\GL(2)$ have
Lie algebras that are abstractly isomorphic, but distinct as subalgebras of
$\gl_2$. Thus $\lie(H)$, seen as an abstract Lie algebra, does not determine
$H$ as a subgroup of $G$.
\end{example}
\autoref{thm:subgroup-lie-alg} fails in positive characteristic. Also,
surjectivity need not hold, see e.g.~$\GL(5)$. If we instead work in the
category of smooth manifolds and Lie groups, then \emph{every} Lie subalgebra
comes from a subgroup.
\begin{theorem}[Cartier]\label{thm:cartier}
Let $k$ be a field of characteristic zero, $G_{/k}$ a linear algebraic group.
Then $G$ is smooth.
\end{theorem}
\begin{proof}
The idea is that for $k=\dC$, $G(\dC)$ must be smooth at ``most'' points
$g$. But any two points in $G(\dC)$ have diffeomorphic neighborhoods, so
$G(\dC)$ itself is smooth. For a careful proof in much greater generality,
see \cite[VI\textsubscript{B} 1.6]{sga3-i}. Even SGA assumes that $G$ is
locally of finite type,
but this is not necessary. See \url{http://mathoverflow.net/questions/22553}
for some discussion and references.
\end{proof}
\begin{example}
Consider $\dmu_p=\ker(\Gm\xrightarrow p \Gm)$ over $k=\dF_p$. That is,
$\dmu_p(A)=\{a\in A:a^p=1\}$. The coordinate ring of $\dmu_p$ is
$k[x]/(x^p-1)$. Note that
\begin{align*}
\lie(\dmu_p)(A)
&= \{1+\varepsilon x:x\in A\text{ and }(1+\varepsilon x)^p=1\} \\
&= \{1+\varepsilon x:1+(\varepsilon x)^p = 1\} \\
&= A .
\end{align*}
In other words, $\dmu_p\subset \GL(1)$ has Lie algebra $\gl_1$. Since $\dmu_p$
is connected over $\dF_p$, this provides a counterexample to
\autoref{thm:subgroup-lie-alg} in characteristic $p$. Moreover,
$\dmu_p(\overline{\dF_p})=1$, so ``$\dmu_p$ has no points.''
\end{example}
If, on the other hand, the base field has
characteristic not $p$, then $(1+\varepsilon x)^p = 1+p\varepsilon x$, which is
$1$ exactly when $x=0$. So $\lie(\dmu_p)=0$ as expected.