-
Notifications
You must be signed in to change notification settings - Fork 0
/
6_special-topics.tex
1263 lines (1074 loc) · 54.5 KB
/
6_special-topics.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
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
% !TEX root = 6490.tex
\section{Special topics}
\subsection[Borel-Weil theorem]{Borel-Weil theorem\footnote{Balazs Elek}}
Let $k$ be a field of characteristic zero, $G_{/k}$ an algebraic group. We
write $\modules(G)$ for the category whose objects are (not necessarily
finite-dimensional) vector spaces $V$ together with $\rho:G\to \GL(V)$. Here,
$\GL(V)$ is the fppf sheaf (not representable unless $V$ is finite-dimensional)
$S\mapsto \GL(V_{\sO(S)})$. The action action of $G$ on itself by
multiplication induces an action of $G$ on the ($k$-vector spaces) $\sO(G)$,
which is not finite-dimensional unless $G$ is finite. By
\cite[I 3.9]{jantzen-2003}, the category $\modules(G)$ has enough injectives;
this enables us to define derived functors in the usual way.
Let $H\subset G$ be an algebraic subgroup. There is an obvious functor
$\restrict_H^G:\modules(G)\to \modules(H)$ which sends $(V,\rho:G\to \GL(V))$
to $(V,\rho|_H)$. It has a right adjoint, the \emph{induction functor},
determined by
\[
\hom_G(V,\induce_H^G U) = \hom_H(\restrict_H^G V,U) .
\]
We are especially interested in this when $U$ is finite-dimensional, in which
case we have
\[
\induce_H^G U = \{f:G\to \dV(U):f(g h) = h^{-1} f(g)\text{ for all }g\in G\} .
\]
Since the induction functor is a right adjoint, it is left-exact, so it
makes sense to talk about its derived functors $\eR^\bullet\induce_H^G$. It
turns out that these can be computed as the sheaf cohomology of certain
locally free sheaves on the quotient $G/H$. Let $\pi:G\epic G/H$ be the
quotient map; for a representation $V$ of $H$, define an $\sO_{G/H}$-module
$\sL(V)$ by
\[
\sL(V)(U) = (V\otimes \sO(\pi^{-1} U))^G .
\]
By \cite[I 5.9]{jantzen-2003}, the functor $\sL(-)$ is exact, and sends
finite-dimensional representations of $H$ to coherent $\sO_{G/H}$-modules.
Moreover, by \cite[I 5.12]{jantzen-2003} there is a canonical isomorphism
\begin{equation*}\tag{$\ast$}\label{eq:induce-cohomology}
\eR^\bullet\induce_H^G V = \h^\bullet(G/H,\sL(V)) .
\]
In general, both the vector spaces in \eqref{eq:induce-cohomology} will not be
finite-dimensional. However, if $G/H$ is proper, then finiteness theorems
for proper pushforward \cite[3.2.1]{ega3-i} tell us that
$\h^\bullet(G/H,\sF)$ is finite-dimensional whenever $\sF$ is coherent. So
we can use $\eR^i \induce_H^G$ to produce finite-dimensional representations of
$G$ from finite-dimensional representations of $H$.
\emph{Note}: for the remainder of this section, ``representation'' means
\emph{finite-dimensional} representation, while ``module'' means possibly
infinite-dimensional representation.
Let $G_{/k}$ be a split reductive group, $B\subset G$ a Borel subgroup and
$T\subset B$ a maximal torus. Let $N=\urad B$; one has $B\simeq T\rtimes N$.
In particular, we can extend $\chi\in \characters^\ast(T)$ to a one-dimensional
representation of $B$ by putting $\chi(t n) = \chi(t)$ for $t\in T$, $n\in N$.
\begin{theorem}
Every irreducible representation of $G$ is a quotient of $\induce_B^G\chi$ for
a unique $\chi\in\characters^\ast(T)$.
\end{theorem}
\begin{proof}
Let $V$ be an irreducible representation of $G$. The group $B$ acts on the
projective variety $\dP(V)$; by \ref{thm:borel-fixed} there is a fixed point
$v$, i.e.~$\restrict_B^G V$ contains a one-dimensional subrepresentation.
This corresponds to a $B$-equivariant map $\chi\monic \restrict_B^G V$ for some
$\chi\in \characters^\ast(T)$. By the definition of induction functors, we get
a nonzero map $\induce_B^G \chi \to V$. Since $V$ is simple, it must be
surjective. Uniqueness of $\chi$ is a bit trickier.
\end{proof}
One calls $\chi$ the \emph{highest weight} of $V$. A natural question is: for
which $\chi$ is $\induce_B^G\chi$ irreducible? By \cite[II 2.3]{jantzen-2003},
$\induce_B^G\chi$ (if nonzero) contains a unique simple subrepresentation,
which we denote by $L(\chi)$.
Recall that our choice of Borel $B\subset T$ induces a base
$S\subset \roots(G,T)\subset \characters^\ast(T)$. We put an ordering on
$\characters^\ast(T)$ by saying that $\lambda \leqslant \mu$ if
$\mu-\lambda\in \dN\cdot S$. It is known that there exists
$w_0\in W=\weyl(G,T)$ such that $w_0(R^+)=R^-$. Finally, the set of
\emph{dominant weights} is:
\[
\characters^\ast(T)_+ = \{\chi\in \characters^\ast(T):\langle \chi,\check\alpha\rangle \geqslant 0\text{ for all }\alpha\in R^+\} .
\]
We can now classify irreducible representations of $G$.
\begin{theorem}
Any irreducible representation of $G$ is of the form $L(\chi)$ for a unique
$\chi\in \characters^\ast(T)_+$.
\end{theorem}
\begin{proof}
This is \cite[II 2.7]{jantzen-2003}.
\end{proof}
The Borel-Weil theorem completely describes $\eR^\bullet\induce_B^G(\chi)$ for
dominant $\chi$. First we need some definitions. If $w\in W$, the \emph{length}
of $w$, denoted $l(w)$, is the minimal $n$ such that $w$ can be written as a
product $s_1\dotsm s_n$ of simple reflections. Let $S$ be a set of simple roots,
$\rho=\frac 1 2 \sum_{\alpha\in R^+} \alpha$. The ``dot action'' of $W$ on
character is:
\[
w\bullet \chi = w(\chi+\rho)-\rho .
\]
Define
\[
C = \{\chi\in \characters^\ast(T):\langle \chi+\rho,\check\alpha\rangle\text{ for all }\alpha\in R^+\} .
\]
It turns out that all $\chi\in \characters^\ast(T)$ are of the form
$w\bullet \chi_1$ for some $\chi_1\in C$. Thus the following theorem describes
$\eR^\bullet\induce_B^G\chi$ for all $\chi$.
\begin{theorem}[Borel-Weil]
Let $\chi\in C$. If $c\notin \characters^\ast(T)_+$, then
$\eR^\bullet\induce_B^G(w\bullet\chi)=0$ for all $w\in W$. If
$\chi\in \characters^\ast(T)_+$, then for all $w\in W$,
\[
\eR^i\induce_B^G(w\bullet\chi) = \begin{cases} L(\chi)& i=l(w) \\ 0 & \text{otherwise} \end{cases}
\]
\end{theorem}
\begin{proof}
This is \cite[II 5.5]{jantzen-2003}.
\end{proof}
Putting $w=1$, we see that $\induce_B^G(\chi) = L(\chi)$.
\subsection{Tannakian categories}
Throughout, $k$ is an arbitrary field of characteristic zero. We will work over
$k$, so all maps are tacitly assumed to be $k$-linear and all tensor product
will be over $k$. Consider the following categories.
For $G_{/k}$ an algebraic group, the category $\rep(G)$ has as objects pairs
$(V,\rho)$, where $V$ is a finite-dimensional $k$-vector space and
$\rho:G\to \GL(V)$ is a homomorphism of $k$-groups. A morphism
$(V_1,\rho_1)\to (V_2,\rho_2)$ in $\rep(G)$ is a $k$-linear map
$f:V_1\to V_2$ such that for all $k$-algebras $A$ and $g\in G(A)$, one has
$f \rho_1(g) = \rho_2(g) f$, i.e.~the following diagram commutes:
\[
\begin{tikzcd}
V_1\otimes A \ar[r, "f"] \ar[d, "\rho_1(g)"]
& V_2\otimes A \ar[d, "\rho_2(g)"] \\
V_1\otimes A \ar[r, "f"]
& V_2\otimes A .
\end{tikzcd}
\]
\begin{example}[Representations of a Hopf algebra]
Let $H$ be a co-commutative Hopf algebra. The category $\rep(H)$ has as objects
$H$-modules that are finite-dimensional over $k$, and morphisms are
$k$-linear maps. The algebra $H$ acts on a tensor product $U\otimes V$ via
its comultiplication $\Delta:H\to H\otimes H$.
\end{example}
\begin{example}[Representations of a Lie algebra]
Let $\fg$ be a Lie algebra over $k$. The category $\rep(\fg)$ has as objects
$\fg$-representations that are finite-dimensional as a $k$-vector space. There
is a canonical isomorphism $\rep(\fg)=\rep(\cU \fg)$, where $\cU \fg$ is the
universal enveloping algebra of $\fg$.
\end{example}
\begin{example}[Continuous representations of a compact Lie group]
Let $K$ be a compact Lie group. The category $\rep_\dC(K)$ has as objects
pairs $(V,\rho)$, where $V$ is a finite-dimensional complex vector space and
$\rho:K\to \GL(V)$ is a continuous (hence smooth, by Cartan's theorem)
homomorphism. Morphisms $(V_1,\rho_1)\to (V_2,\rho_2)$ are $K$-equivariant
$\dC$-linear maps $V_1\to V_2$.
\end{example}
\begin{example}[Graded vector spaces]
Consider the category whose objects are finite-dimensional $k$-vector spaces
$V$ together with a direct sum decomposition $V=\bigoplus_{n\in \dZ} V_n$.
Morphisms $U\to V$ are $k$-linear maps $f:U\to V$ such that
$f(U_n)\subset V_n$.
\end{example}
\begin{example}[Hodge structures]
Let $V$ be a finite-dimensional $\dR$-vector space. A \emph{Hodge structure}
on $V$ is a direct sum decomposition $V_\dC=\bigoplus V_{p,q}$ such that
$\overline{V_{p,q}}=V_{q,p}$. If $U,V$ are vector spaces with Hodge structures,
a morphism $U\to V$ is a $\dR$-linear map $f:U\to V$ such that
$f(U_{p,q})\subset V_{p,q}$. Write $\mathsf{Hdg}$ for the category of vector spaces
with Hodge structure.
\end{example}
Let $\mathsf{Vect}(k)$ be the category of finite-dimensional $k$-vector spaces. For
$\cC$ any of the categories above, there is a faithful functor
$\omega:\cC\to \mathsf{Vect}(k)$. In our examples, it is just the forgetful functor.
The main theorem will be that for $\pi=\automorphisms(\omega)$, the functor $\omega$
induces an equivalence of categories $\cC\iso \rep(\pi)$. We proceed to make
sense of the undefined terms in this theorem.
Our definitions follow \cite{deligne-milne-1982}. As before, $k$ is an
arbitrary field of characteristic zero.
\begin{definition}
A \emph{$k$-linear category} is an abelian category $\cC$ such that each
$V_1,V_2$, the group $\hom(V_1,V_2)$ has the structure of a $k$-vector space
in such a way that the composition map
$\hom(V_2,V_3)\otimes \hom(V_1,V_2)\to \hom(V_1,V_3)$ is $k$-linear. For us,
a \emph{rigid $k$-linear tensor category} is a $k$-linear category $\cC$
together with the following data:
\begin{enumerate}
\item An exact faithful functor $\omega:\cC\to \mathsf{Vect}(k)$.
\item A bi-additive functor $\otimes:\cC\times \cC\to \cC$.
\item Natural isomorphisms
$\omega(V_1\otimes V_2)\iso \omega(V_1)\otimes \omega(V_2)$.
\item Isomorphisms $V_1\otimes V_2\iso V_2\otimes V_1$ for all $V_i\in \cC$.
\item Isomorphisms $(V_1\otimes V_2)\otimes V_3\iso V_1\otimes (V_2\otimes V_3)$
\end{enumerate}
These data are required to satisfy the following conditions:
\begin{enumerate}
\item There exists an object $1\in \cC$ such that $\omega(1)$ is
one-dimensional and such that the natural map $k\to \hom(1,1)$ is an
isomorphism.
\item If $\omega(V)$ is one-dimensional, there exists $V^{-1}\in \cC$ such
that $V\otimes V^{-1}\simeq 1$.
\item Under $\omega$, the isomorphisms 3 and 4 are the obvious ones.
\end{enumerate}
\end{definition}
By \cite[Pr.~1.20]{deligne-milne-1982}, this is equivalent to the standard
(more abstract) definition. Note that all our examples
are rigid $k$-linear tensor categories. One calls the
functor $\omega$ a \emph{fiber functor}.
Let $(\cC,\otimes)$ be a rigid $k$-linear tensor category. In this setting,
define a functor $\automorphisms(\omega)$ from $k$-algebras to groups by setting:
\begin{align*}
\automorphisms^\otimes(\omega)(A)
&= \automorphisms^\otimes\left(\omega:\cC\otimes A\to \rep(A)\right) \\
&= \left\{(g_V)\in \prod_{V\in \cC} \GL(\omega(V)\otimes A):\begin{array}{c}g_{V_1\otimes V_2} = g_{V_1}\otimes g_{V_2}\text{, and } \\ f g_{V_1} = g_{V_1} f\text{ for all }f,V_1,V_2\end{array} \right\} .
\end{align*}
In other words, an element of $\automorphisms(\omega)(A)$ consists of a collection
$(g_V)$ of $A$-linear automorphisms
$g_V:\omega(V)\otimes A\iso \omega(V)\otimes A$, where $V$ ranges over objects
in $\cC$. This collection must satisfy:
\begin{enumerate}
\item $g_1 = 1_{\omega(1)}$
\item $g_{V_1\otimes V_2} = g_{V_1}\otimes g_{V_2}$ for all $V_1,V_2\in \cC$, and
\item whenever $f:V_1\to V_2$ is a morphism in $\cC$, the following diagram
commutes:
\[
\begin{tikzcd}
\omega(V_1)_A \ar[r, "f"] \ar[d, "g_{V_1}"]
& \omega(V_2)_A \ar[d, "g_{V_2}"] \\
\omega(V_1)_A \ar[r, "f"]
& \omega(V_2)_A .
\end{tikzcd}
\]
\end{enumerate}
Typically one only considers affine group schemes $G_{/k}$ that are
\emph{algebraic}, i.e.~whose coordinate ring $\sO(G)$ is a finitely generated
$k$-algebra, or equivalently that admit a finite-dimensional faithful
representation. Let $G_{/k}$ be an arbitrary affine group scheme, $V$ an
arbitrary representation of $G$ over $k$. By
\cite[Cor.~2.4]{deligne-milne-1982}, one has $V=\varinjlim V_i$, where $V_i$
ranges over the finite-dimensional subrepresentations of $V$. Applying this to
the regular representation $G\to \GL(\sO(G))$, we see that
$\sO(G)=\varinjlim\sO(G_i)$, where $G_i$ ranges over the algebraic quotients of
$G$. That is, an arbitrary affine group scheme $G_{/k}$ can be written as a
filtered projective limit $G=\varprojlim G_i$, where each $G_i$ is an affine
algebraic group over $k$. So we will speak of pro-algebraic groups instead of
arbitrary affine group schemes.
If $V$ is a finite-dimensional $k$-vector space and $G=\varprojlim G_i$ is a
pro-algebraic $k$-group, representations $G\to \GL(V)$ factor through some
algebraic quotient $G_i$. That is,
$\hom(G,\GL(V))=\varinjlim \hom(G_i,\GL(V))$. As a basic example of this,
let $\Gamma$ be a profinite group, i.e.~a projective limit of finite groups. If
we think of $\Gamma$ as a pro-algebraic group, then algebraic representations
$\Gamma\to \GL(V)$ are exactly those representations that are continuous when
$V$ is given the discrete topology.
First, suppose $\cC=\rep(G)$ for a pro-algebraic group $G$, and that
$\omega:\rep(G)\to \mathsf{Vect}(k)$ is the forgetful functor. Then the Tannakian
fundamental group $\automorphisms^\otimes(\omega)$ carries no new information
\cite[Pr.~2.8]{deligne-milne-1982}:
\begin{theorem}\label{thm:reconst}
Let $G_{/k}$ be a pro-algebraic group, $\omega:\rep(G)\to \mathsf{Vect}(k)$ the
forgetful functor. Then $G\iso \automorphisms^\otimes(G)$.
\end{theorem}
The main theorem is the following, taken essentially verbatum from
\cite[Th.~2.11]{deligne-milne-1982}.
\begin{theorem}\label{thm:main}
Let $(\cC,\otimes,\omega)$ be a rigid $k$-linear tensor category. Then
$\pi=\automorphisms^\otimes(\omega)$ is represented by a pro-algebraic group, and
$\omega:\cC\to \rep(\pi)$ is an equivalence of categories.
\end{theorem}
Often, the group $\pi_1(\cC)$ is ``too large'' to handle directly. For
example, if $\cC$ contains infinitely many simple objects, probably
$\pi_1(\cC)$ will be infinite-dimensional. For $V\in \cC$, let
$\cC(V)$ be the Tannakian subcategory of $\cC$ generated by $V$. One
puts $\pi_1(\cC/V)=\pi_1(\cC(V))$. It turns out that
$\pi_1(\cC/V)\subset \GL(\omega V)$, so $\pi_1(\cC/V)$ is finite-dimensional.
One has $\pi_1(\cC)=\varprojlim \pi_1(\cC/V)$.
\begin{example}[Pro-algebraic groups]
If $G_{/k}$ is a pro-algebraic group, then \autoref{thm:reconst} tells us that if
$\omega:\rep(G)\to \mathsf{Vect}(k)$ is the forgetful functor, then
$G=\automorphisms^\otimes(G)$. That is, $G=\pi_1(\rep G)$.
\end{example}
\begin{example}[Hopf algebras]
Suppose $H$ is a co-commutative Hopf algebra over $k$. Then
$\pi_1(\rep H)=\spectrum(H^\circ)$, where $H^\circ$ is the \emph{reduced dual}
defined in \cite{cartier-2007}. Namely, for any $k$-algebra $A$, $A^\circ$ is
the set of $k$-linear maps $\lambda:A\to k$ such that $\lambda(\fa)=0$ for some
two-sided ideal $\fa\subset A$ of finite codimension. The key fact here is that
$(A\otimes B)^\circ=A^\circ\otimes B^\circ$, so that we can use multiplication
$m:H\otimes H\to H$ to define comultiplication
$m^\ast:H^\circ\to (H\otimes H)^\circ=H^\circ\otimes H^\circ$.
From \cite[II \S 6 1.1]{demazure-gabriel-1980}, if $G$ is a linear algebraic
group over an algebraically closed field $k$ of characteristic zero, we get an
isomorphism $\sO(G)^\circ= k[G(k)]\otimes \cU(\fg)$. Here $k[G(k)]$ is the
usual group algebra of the abstract group $G(k)$, and $\cU(\fg)$ is the
universal enveloping algebra of $\fg=\lie(G)$, both with their standard Hopf
structures.
\end{example}
[Note: one often calls $\sO(G)^\circ$ the ``space of distributions on $G$.''
If instead $G$ is a real Lie group, then one often writes $\sH(G)$ for the
space of distributions on $G$. Let $K\subset G$ be a maximal compact subgroup,
$M(K)$ the space of finite measures on $K$. Then convolution
$D\otimes \mu\mapsto D\ast\mu$ induces an isomorphism
$\cU(\fg)\otimes M(K)\iso \sH(G)$. In the algebraic setting,
$k[G(k)]$ is the appropriate replacement for $M(K)$.]
\begin{example}[Lie algebras]
Let $\fg$ be a semisimple Lie algebra over $k$. Then by \cite{milne-2007},
$G=\pi_1(\rep\fg)$ is the unique connected, simply connected algebraic group
with $\lie(G)=\fg$. If $\fg$ is not semisimple, e.g.~$\fg=k$, then things get a
lot nastier. See the above example.
\end{example}
\begin{example}[Compact Lie groups]
By definition, the \emph{complexification} of a real Lie group $K$ is a complex
Lie group $K_\dC$ such that all morphisms $K\to \GL(V)$ factor uniquely through
$K_\dC\to\GL(V)$. It turns out that $K_\dC$ is a complex algebraic group, and
so $\pi_1(\rep K)=K_\dC$.
\end{example}
\begin{example}[Graded vector spaces]
To give a grading $V=\bigoplus_{n\in \dZ} V_n$ on a vector space is equivalent
to giving an action of the split rank-one torus $\Gm$. On each $V_n$, $\Gm$
acts via the character $g\mapsto g^n$. Thus
$\pi_1(\text{graded vector spaces})=\Gm$.
\end{example}
\begin{example}[Hodge structures]
Let $\dS=\weil_{\dC/\dR}\Gm$; this is defined by $\dS(A)=(A\otimes\dC)^\times$
for $\dR$-algebras $A$. One can check that the category $\mathsf{Hdg}$ of Hodge
structures is equivalent to $\rep_\dR(\dS)$. Thus $\pi_1(\mathsf{Hdg})=\dS$.
\end{example}
\subsection[Automorphisms of semisimple Lie algebras]{Automorphisms of semisimple Lie algebras\footnote{Sasha Patotski}}
Let $k$ be a field of characteristic zero, $\fg$ a split semisimple Lie algebra
over $k$. We want to describe the group $\automorphisms(\fg)$. It turns out
$\automorphisms(\fg)=\automorphisms(G^\mathrm{sc})$, where $G^\mathrm{sc}$ is
the unique simply connected semisimple group with Lie algebra $\fg$. We can give
$\automorphisms(\fg)$ the structure of a linear algebraic group by putting
\[
\automorphisms(\fg)(A) = \automorphisms(\fg_A) .
\]
Clearly $\automorphisms(\fg)\subset \GL(\fg)$. Let $\ft\subset \fg$ be a maximal
abelian subalgebra. Let $R\subset \ft^\vee$ be the set of roots,
$S\subset R$ a base of $\Delta$. We can define some subgroups of
$\automorphisms(\fg)$:
\begin{align*}
\automorphisms(\fg,\ft)
&= \{\theta\in \automorphisms(\fg):\theta(\ft)= \ft\} \\
\automorphisms(\fg,\ft,S)
&= \{\theta\in \automorphisms(\fg,\ft):\transpose\theta(S)= S\} .
\end{align*}
\begin{lemma}
If $\theta\in \automorphisms(\fg,\ft)$, then $\transpose\theta(R)=R$.
\end{lemma}
\begin{proof}
The definition of $R$ is invariant under automorphisms of $\ft$.
\end{proof}
Let $\automorphisms(R)$ be the set of automorphisms of the Dynkin diagram
associated to $R$.
\begin{lemma}
The natural map
$\varepsilon:\automorphisms(\fg,\ft,S)\to \automorphisms(R)$ is surjective.
\end{lemma}
\begin{proof}
Choose, for each $\alpha\in R$, a non-zero element $x_\alpha\in \fg_\alpha$.
The existence theorem \cite[VIII \S 4.4 th.2]{bourbaki-lie-alg-7-9} tells us
that to each automorphism $\varphi$ of the Dynkin diagram of $R$, there exists
a unique $\tilde\varphi\in \automorphisms(\fg,\ft,S,\{x_\alpha\})$ inducing
$\varphi$. Thus $\varepsilon:\automorphisms(\ft,\ft,S)\to \automorphisms(R)$
splits.
\end{proof}
We want to describe the kernel of $\varepsilon$. Since $\fg$ is semisimple,
the adjoint map $\fg\to \derivations(\fg)=\lie(\automorphisms\fg)$ is an
embedding, and thus $\ft\iso \adjoint(\ft)$. Let $T\subset \inner(\fg)$ be the
subgroup with Lie algebra $\adjoint(\ft)$.
\begin{lemma}
$\ker(\varepsilon) = T$.
\end{lemma}
\begin{proof}
Clearly $T\subset \ker(\varepsilon)$. Let $\theta\in \ker(\varepsilon)$. For
any $\alpha\in R$, we have $\theta x_\alpha\in \fg_\beta$ for some $\beta$.
For $t\in\ft$, we compute:
\begin{align*}
[t,\theta x_\alpha]
&= [\theta t,\theta x_\alpha] \\
&= \theta[t,x_\alpha] \\
&= \alpha(t)\theta x_\alpha ,
\end{align*}
so $\alpha=\beta$. Moreover, $\theta$ acts on $\fg$ exactly like an element of
$\ft$, so $\theta\in T$.
\end{proof}
\begin{lemma}
$\automorphisms(\fg)=\automorphisms(\fg,\ft,S)\cdot \inner(\fg)$.
\end{lemma}
\begin{proof}
Use the well-known facts that all Cartan subalgebras of $\fg$ are conjugate,
and that moreover the Weyl group acts transitively on sets of simple roots.
\end{proof}
We have arrived at the main result.
\begin{theorem}
$\automorphisms(\fg)=\inner(\fg)\rtimes \automorphisms(R)$. In
particular, $\mathrm{Out}(\fg) = \automorphisms(R)$.
\end{theorem}
This allows us to recover the table of automorphisms in
\autoref{sec:root-classify}.
\subsection[Exceptional isomorphisms]{Exceptional isomorphisms\footnote{Theodore Hui}}
Terence Tao's blog post at
\url{http://terrytao.wordpress.com/2011/03/11/}
is an excellent reference for this section. From
\autoref{thm:classify-semisimple}, we know that the set of isogeny classes of
split semisimple algebraic groups is the same as the set of isomorphism classes
of Dynkin diagrams. In \autoref{thm:classify-root}, we classified the Dynkin
diagrams. In this classification, we just included, e.g.~$\typeD_n$ for
$n\geqslant 4$. If we include all the $\typeD_n$ etc., the classification is
no longer unique -- we have to account for the ``exceptional isomorphisms''
\begin{align*}
\typeA_1 &\simeq \typeB_1\simeq \typeC_1\simeq \typeD_2\simeq \typeE_1 \\
\typeB_2 &\simeq \typeC_2 \\
\typeD_3 &\simeq \typeA_3 \\
\typeD_2 &\simeq \typeA_1\times \typeA_1 .
\end{align*}
The Dynkin diagram for $\typeD_2$ and $\typeA_1\times \typeA_1$ is disconnected
-- it is a disjoint union of two points. We'll explicitly construct the induced
isogenies between algebraic groups of different types.
\begin{example}[$\typeA_1\simeq\typeC_1$]
Since $\Sp(2)=\SL(2)$, there is nothing to prove.
\end{example}
\begin{example}[$\typeA_1\simeq\typeB_1$]
Define a pairing on $\Sl_2$ by $\langle x,y\rangle = \trace(x y)$ (the
\emph{Killing form}). One easily verifies that the adjoint action of
$\SL(2)$ on $\Sl_2$ preserves this form. Moreover, a general theorem of linear
algebra tells us that any non-degenerate bilinear symmetric pairing on a
three-dimensional vector space is isomorphic to the orthogonal pairing. It
follows that $\automorphisms(\Sl_2,\langle\cdot,\cdot\rangle)\simeq \Or(3)$,
at least over an algebraically closed field. Since $\SL(2)$ is connected,
$\adjoint(\SL(2))\supset \SO(3)$. A dimension count tells us that
$\SL(2)/\dmu_2\simeq \SO(3)$. Again, this only works over an algebraically
closed field.
\end{example}
\begin{example}[$\typeD_2\simeq \typeA_1\times \typeA_1$]
We need to show that $\SO(4)$ and $\SL(2)\times \SL(2)$ are isogenous.
Let $\mathrm{std}:\SL(2)\monic \GL(2)$ be the standard representation, and
consider the representation $\mathrm{std}\boxtimes\mathrm{std}$ of
$\SL(2)\times \SL(2)$. There is an obvious bilinear form form:
\[
\langle u_1\otimes u_2,v_1\otimes v_2\rangle = \omega(u_1,v_1)\omega(u_2,v_2) ,
\]
where $\omega$ is the determinant pairing
$k^2\times k^2\to \bigwedge^2 k^2\simeq k$ given by $(x,y)\mapsto x\wedge y$.
The group $\SL(2)\times \SL(2)$ acts on $\mathrm{std}\boxtimes\mathrm{std}$ by
$(g,h)(v\otimes w) = (g v)\otimes (g w)$. Thus we have a representation
$\SL(2)\times \SL(2)\to \GL(4)$. The image preserves
$\langle\cdot,\cdot\rangle$, hence (by linear algebra) lies inside $\Or(4)$,
By connectedness and a dimension count, we see that this is an isogeny
$\SL(2)\times \SL(2)\epic \SO(4)$.
\end{example}
\subsection[Constructing some exceptional groups]
{Constructing some exceptional groups\footnote{Gautam Gopal}}
We roughly follow \cite{springer-veldkamp-2000}. Let $k$ be a field of
characteristic not $2$ or $3$. Recall that if $V$ is a $k$-vector space and
$\langle\cdot,\cdot\rangle:V\times V\to k$ is a symmetric bilinear form, we can
define a quadratic form $q:V\to k$ by $q(v)=\langle v,v\rangle$. This
correspondence is bijective; we can go backwards via the familiar identity
\[
\langle u,v\rangle = \frac 1 2 (q(u+v)-q(u)-q(v)) .
\]
\begin{definition}
A \emph{composition algebra} is a pair $(C,q)$, where $C$ is a unital,
not-necessarily associative $k$-algebra and $q$ is a multiplicative
non-degenerate quadratic on $C$.
\end{definition}
In other words, we require $q(x y)=q(x)q(y)$ for all $x,y\in C$. Since by
\cite[1.2.4]{springer-veldkamp-2000}, $q$ is determined by the multiplicative
strucure of $C$, we will just refer to ``a composition algebra $C$.'' Write $e$
for the unit of $C$. Every composition algebra comes with a natural involution
$x\mapsto \bar x$, defined by $\bar x=\langle x,e\rangle-x$.
\begin{theorem}
Let $C$ be a composition algebra. Then $\dim(C)\in \{1,2,4,8\}$.
\end{theorem}
\begin{proof}
This is \cite[1.6.2]{springer-veldkamp-2000}.
\end{proof}
We call an $8$-dimensional composition algebra an \emph{octobian algebra}.
If $C$ is an octonian $k$-algebra, we define an algebraic group
$\automorphisms(C)$ by putting
\[
\automorphisms(C)(A) = \{g\in \GL(C\otimes A):g\text{ is a morphism of normed $A$-algebras}\} ,
\]
for all (commutative, unital) $k$-algebras $A$. There is an obvious embedding
$\automorphisms(C)\monic \Or(C,q)$.
\begin{theorem}
Let $C$ be an octonian $k$-algebra. Then $\automorphisms(C)$ is a connected
algebraic group of type $\typeG_2$.
\end{theorem}
\begin{proof}
We mean that after base-change to an algebraic closure of $k$,
$\automorphisms(C)$ becomes isomorphic to $\typeG_2$. This is
\cite[2.3.5]{springer-veldkamp-2000}.
\end{proof}
If $C$ is a composition algebra and
$\gamma=(\gamma_1,\gamma_2,\gamma_3)\in (k^\times)^3$, we define a new algebra
$H_{C,\gamma}$ to be as a set the collection of matrices
\[
\begin{pmatrix}
z_1 & c_3 & \gamma_1^{-1} \gamma_3 \bar c_2 \\
\gamma_2^{-1} \gamma_1 \bar c_3 & z_2 & c_1 \\
c_2 & \gamma_3^{-1} \gamma_2 \bar c_1 & z_3 \end{pmatrix} \qquad c_i\in k^\times\text{ and }z_i\in k .
\]
Give $H_{C,\gamma}$ the product $x y = \frac 1 2 (x\cdot y+y\cdot x)$ and
quadratic form $q(x)=\frac 1 2 \trace(x^2)$.
\begin{definition}
An \emph{Albert algebra} is a commutative, non-unital, associative $k$-algebra
$A$ such that $A\otimes \bar k$ is isomorphic to a $\bar k$-algebra of the form
$H_{C,\gamma}$ for some octobian algebra $C$ and $\gamma\in \bar k^\times$.
\end{definition}
If $A$ is an Albert algebra, we can define an algebraic group
$\automorphisms(A)$ just as above.
\begin{theorem}
If $A$ is an Albert algebra, then $\automorphisms(A)$ is a connected simple
algebraic group of type $\typeF_4$.
\end{theorem}
\begin{proof}
Put $G=\automorphisms(A)$.
Call an element $u\in A$ \emph{idempotent} if $u2=u$. It turns out that if $u$
is idempotent, then either $u\in \{0,e\}$, or $q(u)\in \{1/2,1\}$. Call the
idempotents with $q(u)=1/2$ \emph{primitive}. Let $V\subset A$ be the set of
primitive idempotents; this is naturally a variety over $k$. It turns out that
$V$ is $16$-dimensional closed and irreducible and has a transitive $G$-action.
For some $v\in V$, the group $G_v=\stabilizer_G(v)$ is the spin group of a
nine-dimensional quadratic form, so $\dim(G_v)=36$. It follows
that $\dim(G)=16+36=52$. Since $G_v$ and $V$ are irreducible, $G$ is connected.
From the action of $G$ on $e^\bot\subset A$, we see that $G$ is semisimple
algebraic. The only $52$-dimensional semisimple algebraic group over an
algebraically closed field is $\typeF_4$. For a more careful proof, see
\cite[7.2.1]{springer-veldkamp-2000}.
\end{proof}
It turns out that any group of type $\typeF_4$ can be obtained as
$\automorphisms(A)$ for some Albert algebra $A$. Consider the cubic form
\[
\det(x) = z_1 z_2 z_3-\gamma_3^{-1} \gamma_2 z_1 q(c_1) - \gamma_2^{-1} \gamma_3 z_2 q(c_2) - \gamma_2^{-1} \gamma_1 z_3 q(c_3)+\langle c_1 c_2,\bar c_3\rangle .
\]
Let $\GL(A,\det)$ be the subgroup of $\GL(A)$ consisting of those linear
maps which preserve $\det$.
\begin{theorem}
Let $A$ be an Albert algebra. Then $\GL(A,\det)$ is a connected simple
algebraic group of type $\typeE_6$.
\end{theorem}
\begin{proof}
This is \cite[7.3.2]{springer-veldkamp-2000}.
\end{proof}
Unlike the case with groups of type $\typeF_4$, not all groups of type
$\typeE_6$ can be obtained this way.
\subsection[Spin groups]{Spin groups\footnote{Benjamin ?}}
The motivation for spin groups is as follows. Recall that up to isogeny, the
simple algebraic groups are $\SL(n)$, $\Sp(2n)$, $\SO(n)$, or one of the
exceptional groups. Recall that each isogeny class has two distinguished
elements, the simply connected and adjoint. For $\SO(n)$, we should expect
there to be a simply connected group $\Spin(n)=(\typeD_n)^\mathrm{sc}$, which
is a double cover of $\SO(n)$.
Work over a field $k$ of characteristic not $2$. Let $V$ be a $k$-vector space,
$q$ a quadratic form on $V$. The \emph{Clifford algebra} $\clifford(V,q)$ is
the quotient of the tensor algebra $T(V)$ by the ideal generated by
$\{v\otimes v-q(v):v\in V\}$. There is an obvious injection
$V\monic \clifford(V,q)$, and $k$-linear maps $f:V\to A$ into associative
$k$-algebras lift to $\tilde f:\clifford(V,q)\to A$ if and only if
$f(v) f(v)=q(v)$ for all $v\in V$. This univeral property clearly
characterizes $\clifford(V,q)$. For brevity, write
$\clifford(V)=\clifford(V,q)$.
\begin{example}
Let $V=\dR^4$, $q$ be the indefinite form of signature $(3,1)$,
i.e.~$q(v)=-v_1^2+v_2^2+v3^2+v_4^2$. We write $\clifford_{(1,3)}(\dR)$ for
the Clifford algebra $\clifford(V,q)$; it has presentation
\[
\clifford_{(1,3)}(\dR) = \dR\langle e_1,e_2,e_3,e_4\rangle / (e_1^2=-1, e_2^2=e_3^2=e_4^2=1) .
\]
This is used in the Dirac equation, which unifies special relativity and
quantum mechanics.
\end{example}
There is a clear action $\Or(V,q)\to \automorphisms \clifford(V)$. In
particular, the involution $\alpha(v)=-v$ induces an involution (also denoted
$\alpha$) of $\clifford(V)$. This induces a grading
$\clifford(V)=\clifford^0(V)\oplus \clifford^1(V)$, where
\begin{align*}
\clifford^0(V) &= \{x\in\clifford(V):\alpha(x)=x\} \\
\clifford^1(V) &= \{x\in \clifford(V):\alpha(x)=-x\} .
\end{align*}
We define some algebraic groups via their functors of points:
\begin{align*}
\Pin(V,q)(A) &= \{g\in \clifford(V,q)_A:q(g)\in \dmu_2(A)\} \\
\Spin(V,q) &= \Pin(V,q)\cap \clifford^0(V) .
\end{align*}
There is a ``twisted adjoint map'' $\tilde\adjoint:\Pin(V,q)\to \Or(V,q)$,
given by $\tilde\adjoint(g)(v) = \alpha(g) v g^{-1}$.
\begin{theorem}
There is a natural exact sequence
\[
1 \to \dmu_2 \to \Spin(V,q) \to \SO(V,q) \to 1 .
\]
\end{theorem}
\begin{proof}
This is \cite[IV.10.21]{berhuy-2010}.
\end{proof}
\subsection[Differential Galois theory]
{Differential Galois theory\footnote{Ian Pendleton}}
A good source for differential Galois theory is \cite{vanderput-singer-2003}.
Recall that if $A$ is a ring, a \emph{derivation} on $A$ is an additive map
$\partial:A\to A$ satisfying the \emph{Liebniz rule}:
$\partial(a b)=a\partial(b)+\partial(a)b$. We write $\derivations(A)$ for the
group of derivations $A\to A$.
\begin{definition}
A \emph{differential ring} is a pair $(R,\Delta)$, where
$\Delta\subset \derivations(R)$ is such that
$\partial_1\partial_2 = \partial_2\partial_1$ for all
$\partial_1,\partial_2\in \Delta$.
\end{definition}
If $\Delta=\{\partial\}$, we write $r'=\partial r$ for $r\in R$. The ring
$C=\{c\in R:\partial c=0\text{ for all }\partial\in\Delta\}$ is called the
\emph{ring of constants}. If $R$ is a field, we call $(R,\Delta)$ a
\emph{differential field}.
\begin{example}
Let $R=C^\infty(\dR^n)$ and
$\Delta=\{\frac{\partial}{\partial x_i}:1\leqslant i\leqslant n\}$. Then
$(R,\Delta)$ is a differential ring with $\dR$ as ring of constants.
\end{example}
\begin{example}
Let $k$ be a field, $R=k(x_1,\dots,x_n)$, and
$\Delta=\{\frac{\partial}{\partial x_i}:1\leqslant i\leqslant n\}$. Then
$(R,\Delta)$ is a differential field with field of constants $k$.
\end{example}
We are interested in solving \emph{matrix differential equations}, that is,
equations of the form $y'=A y$ for $A\in \matrices_n(k)$. A solution would be a
tuple $y=(y_1,\dots,y_n)\in k^n$ such that $(y_1',\dots,y_n') = A y$. To a
matrix differential equation we will associate two objects: a Picard-Vessiot ring
$R$, and a linear algebraic group $\dgalois(R/k)$.
\begin{definition}
Let $(k,\partial)$ be a differential field, $(R,\partial)$ a differential
$k$-algebra (so $\partial_R|_k=\partial_k$). Let $A\in \matrices_n(k)$. A
\emph{fundamental solution matrix} to the equation $y'=A y$ is an element
$Z\in \GL_n(R)$ such that $Z'=A Z$.
\end{definition}
It is easy to construct a (universal) fundamental solution matrix for the
equation $y'=A y$. Let $S=\sO(\GL_n)=k[y_{i j},\det(y_{i j})^{-1}]$, and
define a differential $\partial:S\to S$ by $\partial(y_{i j}) = (A y)_{i j}$.
It is easy to see that $S$ represents the functor that sends a differential
$k$-algebra $R$ to the set of fundamental solution matrices in $R$.
If $(R,\Delta)$ is a differential ring, a \emph{differential ideal} is an
ideal $\fa\subset R$ such that $\partial(\fa)\subset \fa$ for all
$\partial\in \Delta$. If $\fa$ is a differential ideal, then $R/\fa$ naturally
has the structure of a differential ring. Call a differential ring
\emph{simple} if it has no nontrivial differential ideals. Note that simple
differential rings need not be fields,
e.g.~$(k[t],\frac{\partial}{\partial t})$.
\begin{definition}
Let $A\in \matrices_n(k)$. A \emph{Picard-Vessiot ring} for the equation
$y'=A y$ is pair $(R,Z)$, where $R$ is a simple differential $k$-algebra which
has a fundamental solution matrix $Z$ for $y'=A y$, such that $R$ is generated
as a $k$-algebra by the entries of $Z$ and $\frac{1}{\det Z}$.
\end{definition}
It is easy to prove that Picard-Vessiot rings exist. Let $S$ be the
differential ring constructed above. For any maximal differential ideal
$\fm\subset S$, the quotient $R=S/\fm$ is a Picard-Vessiot ring. By
\cite[1.20]{vanderput-singer-2003}, Picard-Vessiot rings (for a given
equation $y'=A y$) are unique.
\begin{definition}
Let $(k,\partial)$ be a differential field, $A\in \matrices_n(k)$. The
\emph{differential Galois group} of the equation $y'= A y$ is
$\dgalois(R/k)=\automorphisms_{(k,\partial})(R)$ for any Picard-Vessiot ring
$R$ (for the equation $y'=A y$).
\end{definition}
\begin{lemma}
The group $\dgalois(R/k)$ is linear algebraic.
\end{lemma}
\begin{proof}
See \cite[1.26]{vanderput-singer-2003}.
\end{proof}
\begin{example}
If $k=\dC(x)$ and we consider the equation $y'=\frac{\alpha}{x} y$ for
$\alpha=\frac n m\in \dQ$, then the Picard-Vessiot ring is
$R=\dC(x^{n/m})$ and $\dgalois(R/k)=\dZ/m$.
\end{example}
It is shown in \cite{tretkoff-tretkoff-1979} that over $\dC$, all linear
algebraic groups arise as differential Galois groups. The ``modern'' approach
to differential Galois theory uses $\mathscr{D}$-modules and Tannakian
categories.
\subsection[Universal enveloping algebras and the Poincar\'e-Birkhoff-Witt theorem]
{Universal enveloping algebras and the Poincar\'e-Birkhoff-Witt theorem\footnote{Daoji Huang}}
Let $k$ be a field, $\mathsf{Lie}$ be the category of Lie algebras over $k$,
and $\mathsf{Ass}$ be the category of unital associative $k$-algebras. There
is an easy functor $\cL:\mathsf{Ass}\to \mathsf{Lie}$, that sends a $k$-algebra
$A$ to the Lie algebra $\cL A$ whose underlying vector space is $A$, with
bracket
\[
[a,b] = a\cdot b-b\cdot a .
\]
This has a left adjoint, denoted $\cU$. That is, for each Lie algebra $\fg$,
there is an associative algebra $\cU\fg$ with a $k$-linear map
$i:\fg\to \cU\fg$ satisfying $i[x,y]=[i(x),i(y)]$, such that for any
algebra $A$ and linear map $f:\fg\to A$ satisfying $f[x,y]=[f(x),f(y)]$, there
is a unique extension $\tilde f:\cU\fg\to A$ such that $f=\tilde f\circ i$.
We can construct the universal enveloping algebra $\cU\fg$ directly. Let
$\cT\fg=\bigoplus_{n\geqslant 0} \fg^{\otimes n}$ be the tensor algebra of
$\fg$. It is easy to see that $\cU\fg$ is the quotient of $\cT\fg$ by the
relations $\{x\otimes y-y\otimes x - [x,y]:x,y\in \fg\}$.
There is an obvious filtration on $\cT\fg$, for which
$\cT_m\fg=\bigoplus_{m\geqslant n} \fg^{\otimes m}$. It induces a filtration on
$\cU\fg$. The Poincar\'e-Birkhoff-Witt theorem is an explicit description of
the graded ring $\graded\cU(\fg)$.
Define a map $\varphi:\cT\fg\to \graded(\cU\fg)$ by the obvious surjections
\[
\cT^m\fg = \fg^{\otimes m} \to \cU^m\fg \epic \graded^m\cU(\fg) .
\]
This is a homomorphism of graded $k$-algebras. Moreover, since, for
$x\in \cU_2\fg$, we have
\[
U_2\ni x \otimes y - y\otimes x = [x,y]\in U_1 ,
\]
it follows that the map $\varphi$ factors through the symmetric algebra
$\cS\fg$.
\begin{theorem}[Poincar\'e-Birkhoff-Witt]
Let $\fg$ be a Lie algebra over $k$. The map
$\varphi:\cS(\fg)\to \graded\cU(\fg)$ is an isomorphism of graded $k$-agebras.
\end{theorem}
\begin{proof}
This is \cite[I \S 2.7 th.1]{bourbaki-lie-alg-1-3}.
\end{proof}
More concretely, suppose $\fg$ has a basis $x_1,\dots,x_n$ of $\fg$. The
PBW theorem tells us that every element of $\cU\fg$ can be written uniquely as
a sum
\[
\sum_{\bm i} \lambda_{\bm i} x^{e_{\bm i}} ,
\]
where $\bm i$ ranges over all tuples $(i_1,\dots,i_r)$ for which
$1\leqslant i_1< \cdots < i_r\leqslant n$. Here we write
\[
x^{e_{\bm i}} = x_{i_1}^{e_1} \dotsm x_{i_r}^{e_r} .
\]
The PBW theorem is used in many places -- one application is a simpler (though
less transparent) construction of free Lie algebras than the one in
\cite[II \S 2]{bourbaki-lie-alg-1-3}.
\subsection[Forms of algebraic groups]{Forms of algebraic groups\footnote{Tao Ran Chen}}
We start by defining non-abelian cohomology, following
\cite[II.3]{berhuy-2010}. Let $\Gamma$ be a profinite group, $G$ a (possibly
nonabelian) discrete group on which $G$ acts continuously by automorphisms. The
pointed set of \emph{$1$-cocycles}, denoted $Z^1(\Gamma,G)$, consists of
functions $c:\Gamma\to G$, written $\sigma\mapsto c_\sigma$, such that
\[
c_{\sigma\tau} = c_\sigma \sigma(c_\tau)
\]
for all $\sigma,\tau\in \Gamma$. The distinguished $1$-cocycle is
$\sigma\mapsto 1$. We put an equivalence relation on $Z^1(\Gamma,G)$, namely
$c\sim c'$ if there exists $g\in G$ such that
\[
c'_\sigma = g c_\sigma \sigma(g^{-1})
\]
for all $\sigma\in \Gamma$. Write $\h^1(\Gamma,G)=Z^1(\Gamma,G)/\sim$ be the
set of equivalence classes of $1$-cocycles.
Let $X$ be a $\Gamma$-set with (equivariant) $G$-action, such that the map
$G\times X\to X\times X$ given by $(g,x)\mapsto (g x,x)$ is a bijection. (Such
$X$ are called \emph{$G$-torsors} in the category of $\Gamma$-sets.) Choose
$x_0\in X$ and define $c:\Gamma\to G$ by
$\sigma(x_0) = c_\sigma^{-1}\cdot x_0$. One can check that $c_\sigma$ is a
$1$-cocycle that only depends on the isomorphism class of $X$ as a
$\Gamma$-equivariant $G$-set. The ``right perspective'' here is that $G$ is a
group object in the category of $\Gamma$-sets, and that $\h^1(\Gamma,G)$
classifies $G$-torsors in that category. More generally, let $G$ be a group
object in an arbitrary topos $\cT$. A \emph{$G$-torsor} is a $G$-object
$X\in\cT$ such that $X\to 1$ is an isomorphism, and such that the obvious
morphism $G\times X\to X\times X$ is an isomorphism. As in
\cite[\S 8.3]{johnstone-1977}, one can \emph{define} $\h^1(\cT,G)$ to be the
(pointed) set of isomorphism classes of $G$-torsors in $\cT$. If
$N\subset G$ is a normal subgroup in $\cT$, there is a ``short long exact
sequence'' of pointed sets:
\begin{equation*}
\begin{tikzcd}
1 \ar[r]
& \h^0(\cT,N) \ar[r]
& \h^0(\cT,G) \ar[r] \ar[draw=none]{d}[name=X, anchor=center]{}
& \h^0(\cT,G/N) \ar[rounded corners,
to path={ -- ([xshift=2ex]\tikztostart.east)
|- (X.center) \tikztonodes
-| ([xshift=-2ex]\tikztotarget.west)
-- (\tikztotarget)}]{dll}[at end]{} \\
& \h^1(\cT,N) \ar[r]
& \h^1(\cT,G) \ar[r]
& \h^1(\cT,G/N) .
\end{tikzcd}
\end{equation*}
Our main example is: $\Gamma=\galois(\bar k/k)$ for $k$ a field, in which
case the category of $\Gamma$-sets is equivalent to $\sheaves(k_\etale)$. We
will write $\h^1(k,G)$ instead of $\h^1(k_\etale,G)$ for
$G\in \sheaves(k_\etale)$. Suppose $X$ is an ``object we want to classify.''
Let $G=\automorphisms X$ (internal automorphisms object in the category
$\sheaves(k_\etale)$). If $X'$ is another object and
$\varphi:X'_{k^\mathrm{sep}}\to X_{k^\mathrm{sep}}$ is an isomorphism, we can
define a cocycle $c:\Gamma=\galois(k^\mathrm{sep}/k)\to G(k^\mathrm{sep})$ by
$c_\sigma = \sigma(\varphi)\circ \varphi^{-1}$. This gives a map
$\{k$-forms of $X\}\to \h^1(k,G)$, which will generally be injective. We are
interested in the case where $X$ is an algebraic group over $k$.
Let $k$ be a field, $G_{/k}$ an algebraic group. A \emph{$k$-form of $G$} is
an algebraic group $G'_{/k}$ such that
$G'_{k^\mathrm{sep}} \simeq G_{k^\mathrm{sep}}$. Two $k$-forms $G',G''$ are
equivalent if $G'\simeq G''$ over $k$.
\begin{example}
Let $K/k$ be a finite Galois extension, $G_{/k}$ an algebraic group. We defined
in \autoref{eg:nonsplit-torus} the \emph{Weil restriction} of an algebraic
group; here we apply it to $G_K$. That is, we write $\weil_{K/k} G$ for the
algebraic group over $k$ defined by
\[
(\weil_{K/k} G)(A) = G(A\otimes_k K) ,
\]
for any $k$-algebra $A$. By \cite[\S 7.6]{bosch-lutkebohmert-raynaud-1990},
$\weil_{K/k} G$ is actually an algebraic group. Since
$K\otimes_k K\simeq \prod_\Gamma K$, where $\Gamma=\galois(K/k)$, we have
$(\weil_{K/k} G)_K\simeq \prod_\Gamma G_K$, i.e.~$\weil_{K/k} G$ is a
$k$-form of $\prod_K \Gamma$. Interesting $k$-forms of $G$ can often be
obtained by taking appropriate subgroups of $\weil_{K/k} G$.
\end{example}
\begin{example}
Apply the previous general construction to the extension $\dC/\dR$.
The group $\dS=\weil_{\dC/\dR}\Gm$ comes with a \emph{norm map}
$\norm:\dS\to \Gm$, coming from the ``usual norm''
$\norm:\dC\to \dR$. Then $\dS^{\norm=1}$ is a $\dR$-form of $\Gm$.
\end{example}
\begin{example}
Let $k$ be a field. A \emph{central simple algebra} over $k$ is a (possibly
non-commutative) $k$-algebra $R$ that is simple and has $\zentrum(R)=k$.
The obvious examples are $\matrices_n(k)$, but there are others, for example
any division algebra $D$ with $\zentrum(D)=k$. If $R$ is any central simple
algebra, then $R_{k^\mathrm{sep}}\simeq \matrices_n(k^\mathrm{sep})$ for some
$n$. There is a norm map $\norm:R\to k$, which we can use to define a group
$\SL(R)$ by
\[
\SL(R)(A) = \ker\left((R\otimes_k A)^\times \xrightarrow{\norm} A^\times\right) ,
\]
for all $k$-algebras $A$. Since $R$ splits over $k^\mathrm{sep}$, we have
$\SL(R)_{k^\mathrm{sep}}\simeq \SL(n)_{k^\mathrm{sep}}$.
\end{example}
When $k=\dR$, it is well-known that there is only one division algebra,
the quaternions $\dH$. So the only obvious forms of $\SL(2)_\dR$ are
$\SL(2)$ and $\SL(\dH)$. It turns out that these are the only ones, but this
is far from obvious.
It turns out that the set of $k$-forms of an algebraic group $G$ can be
described using Galois cohomology. Write $\automorphisms(G)$ for the
functor on $k$-schemes defined by
\[
\automorphisms(G)(S) = \automorphisms_S(G_S) .
\]
This is a sheaf for the canonical topology. In particular, we can take
$\h^1(k,\automorphisms G)=\h^1(\galois(\bar k/k),\automorphisms_{k^\mathrm{sep}}(G))$.
Let $G'$ be a $k$-form of $G$. Choose an isomorphism
$\phi:G'_{k^\mathrm{sep}}\iso G_{k^\mathrm{sep}}$, and define
$c:\galois(k^\mathrm{sep}/k)\to \automorphisms_{k^\mathrm{sep}}(G)$ by
$c_\sigma = \sigma(\phi)\circ \phi^{-1}$. It turns out that the class of
$c$ in $\h^1(k,\automorphisms G)$ does not depend on the choice of
$\phi$. In fact, we have the following result.
\begin{theorem}
Let $k$ be a field, $G_{/k}$ an algebraic group. Then the correspondence
defined above induces a natural bijection between the set of equivalence
classes of $k$-forms of $G$ and the pointed set $\h^1(k,\automorphisms G)$.
\end{theorem}