summaryrefslogtreecommitdiff
path: root/web/reduce/rweb/appl/symmetry.tex
blob: eb79fe7509819426d860571230a2bb056da0effb (plain)
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
\documentstyle[a4wide,fleqn]{article}

\input mssymb

\renewcommand{\theequation}{\thesection.\arabic{equation}}
\setlength{\parindent}{0pt}
\newtheorem{df}{Definition}[section]
\newtheorem{lemma}[df]{Lemma}
\newtheorem{th}[df]{Theorem}
\newtheorem{rem}[df]{Remark}
%\newtheorem{proof}[df]{Proof}
\newtheorem{prop}[df]{Proposition}
\newtheorem{cor}[df]{Corollary}
\newtheorem{ex}[df]{Example}
\newcommand{\Ng}{\setcounter{df}{0}}
\newcommand{\rbox}{\begin{flushright} $ \Box $ \end{flushright}}

\newcommand{\groot}{\displaystyle}
\newcommand{\skipline}{\vspace{4mm}}

\def\h{\hbox{$\goth h$}}
\def\g{\hbox{$\goth g$}}
\newtheorem{Definition}{Definition}[section]
\newtheorem{Proposition}[Definition]{Proposition}

\begin{document}

\title{Lecture\\ Generalized Symmetries} \author{Paul H.M. Kersten \\
Department of Applied Mathematics\\ University of Twente\\
P.O. Box 217\\ 7500 AE Enschede\\ The Netherlands}


\date{}
\maketitle

\begin{abstract}
\mbox{\ }
\end{abstract}

\section{Introduction.}
The classical notion of symmetry of a system of differential equations
was based on transformations in the space of independent en dependent
variables, transforming solutions into solutions. These symmetries are
called {\it point} symmetries. The first generalization of this
concept is to consider transformations of independent, dependent
variables and first order partial derivatives, and transforming
solutions into solutions. This leads to the socalled {\it contact}
symmetries. Generalized symmetries, the subject of this lecture, can
be understood as transformations in the space of independent,
dependent variables and {\it all} partial derivatives \cite{O,V}.\\

Notations will be as follows.\\
$X$ is the space of independent variables, local coordinates being
\begin{displaymath}
 (x_1,...,x_p)
\end{displaymath}
$U$ is the space of dependent variables where local coordinates are
\begin{displaymath}
 (u^1,...,u^q)
\end{displaymath}
The $k^{th}$ order jetbundle $J^k(x,u)$ has local coordinates
\begin{equation}
\label{1.1a}
 (x_i,u^{\alpha},u^{\alpha}_I) \qquad (|I| \leq k,i=1,...,p ; \alpha = 1,...,q)
\end{equation}
while the infinite jetbundle $J(x,u) = J^{\infty}(x,u)$ has local
coordinates
\begin{equation}
\label{1.1b}
 (x_i,u^{\alpha},u_I ^{\alpha}) \; \; |I| < \infty
\end{equation}
In (\ref{1.1a}),(\ref{1.1b}) we used the multiindex notation
$I=(i_1,...,i_p) \; |I| = \sum\limits_{k=1}^p i_r$\\

Throughout we shall use summation convention in case an index occurs
twice; latin indices run from $1$ to $p$ while greek indices run from
$1$ to $q$.\\

Functions $f:J^k(x,u) \rightarrow \Bbb R$ are supposed to be
$C^{\infty}$, while functions $g:J(x,u) \rightarrow \Bbb R$ are just
those dependent on a {\it finite} number of variables, so in effect
\begin{displaymath}
 g = \pi_k^* f \mbox{ for some } f \mbox{ and } k,
\end{displaymath}
(see previous lectures).
notation $f = f[u], g = g[u]$.\\
A system of $k-th$ order differential equations is denoted by
\begin{equation}
\label{1.1c}
 \Delta_j[u] = 0 \; (j=1,...,\l)
\end{equation}
where $\Delta_j$ is defined on $J^k(x,u)$.\\
The total partial derivative operators $D_i$ are given by
\begin{equation}
\label{1.1d}
 D_i = \frac{\partial}{\partial x _ {i}} + u^\alpha_{I,i}
\frac{\partial}{\partial u_I^\alpha} \; \; (i=1,...,q) 
\end{equation}
and they re-create in an algebraic way, what is realized classically
by partial differentiation, using chain-rule.\\

In section 2 we give a short recapitulation of the notion of
infinitesimal symmetry. In section 3 the concept of generalized
symmetry is given, some theorems are proved and an explicit example is
given.\\
In section 4 the notion of nonlocal symmetry, \cite{KV,KV2} being a
generalization of generalized symmetry, is introduced and an illustration
through the famous Korteweg-de Vries equation (KdV) is
discussed. In the conclusions we point out that even generalizations
of this concept are very interesting.\\ 

Applications of symmetries to construct explicit solutions,
conservation laws etc are beyond the scope of this lecture, and are
dealt with in p.e. ref \cite{O}, \cite{KB}.

\setcounter{equation}{0}
\section{Classical Symmetries.}
We give a short review of classical (infinitesimal) symmetries of
differential equations.\\
We start at a $k$-th order system of differential equations
\begin{equation}
\label{1.1}
 \Delta_j[u] = 0 \; \; j = 1,\ldots,\l.
\end{equation}
A vector field $V \epsilon T(J^0(x,u))$ is given by
\begin{equation}
\label{1.2}
 V = \xi^i(x,u) \frac{\partial}{\partial x ^ {i}} +
\varphi_{\alpha}(x,u) \frac{\partial}{\partial u ^ {\alpha}}
\end{equation}
The $k^{th}$ prolongation of the vector field $V$ defined in
$T(J^k(x,u))$ and denoted $pr^{k}(V)$ is given by
\begin{equation}
\label{1.3}
 pr^{k}(V) = \xi^i (x,u) \frac{\partial}{\partial x ^ {i}} +
\Phi_{\alpha}^I [u] \frac{\partial}{\partial u
_{I} ^ {\alpha}}
\end{equation}
where
\begin{equation}
\label{1.4}
 \Phi_{\alpha}^I [u] = D^I (\varphi_{\alpha}(x,u) - u_i^{\alpha}
\xi^i (x,u)) + u_{I,i} ^{\alpha} \xi^i (x,u)
\end{equation}
and
\begin{equation}
\label{1.4a}
 D^I = D_1^{i _{1}} o D_2^{i _ {2}} ... o D_p^{i _ {p}}
\end{equation}

Formula (\ref{1.4}) can be obtained by the conditions that the
prolongation of the vector field $V$ leaves the contact structure
\begin{equation}
\label{1.5}
 \omega_J^{\alpha} = du_J^{\alpha} - u_{J,i}^\alpha dx^i \; \;
(|J| \leq k-1)
\end{equation}
invariant \cite{O}.\\
We now arrive at the following definition.

\begin{df}
A vector field $V$ (\ref{1.2}) is a (infinitesimal) symmetry of the
system of differential equations (\ref{1.1}) if
\begin{equation}
\label{1.6}
 \hspace{5cm} pr^{(k)}(V)(\Delta_j) = 0 \; \; \mbox{ on } \Delta = 0
\end{equation}
We shall not compute symmetries here; but postpone it to the next
section.\\
Computerprograms to construct solutions of the symmetry condition
(\ref{1.6}) are discussed in p.e. \cite{K}.
\end{df}

\setcounter{equation}{0}
\section{Generalized Symmetries.}

In this section we generalize the classical notion of infinitesimal
symmetries to generalized symmetries, sometimes called
Lie-B\"{a}cklund transformations: not te be confused with B\"{a}cklund
transformations which are of a completely different nature.\\
Remind that classically a vector field $V \epsilon T(J^0(x,u))$ is
given by
\begin{equation}
\label{2.1}
 V = \xi^i(x,u) \frac{\partial}{\partial x^{i}} + \varphi_\alpha(x,u)
\frac{\partial}{\partial u_{\alpha}}
\end{equation}
We now pass to the infinite jetbundle $J(x,u)$ where local coordinates
are given by
\begin{displaymath}
 (x^i,u^{\alpha},u^{\alpha}_I) \; \; I = (i_1,...,i_p) \; i_k \geq 0(k=1,...,p)
\end{displaymath}
functions $F:J(x,u) \rightarrow \Bbb R$ are to be understood to depend
on an arbitrary but finite number of variables, $F=F[u]$.

\begin{df}
A (formal) generalized vector field is given by the following
expression
\begin{equation}
\label{2.2}
 V = \xi^i[u] \frac{\partial}{\partial x ^ {i}} +
\varphi_{\alpha}[u] \frac{\partial}{\partial u ^ {\alpha}}
\end{equation}
The formal prolongation of $V$ to the infinite jetbundle is defined by
\begin{equation}
\label{2.3}
 pr(V) = \xi^i[u] \frac{\partial}{\partial x ^{i}} +
\Phi^J_{\alpha}[u] \frac{\partial}{\partial u^{\alpha}_{J}}
\end{equation}
whereas in the second term summation runs over $\alpha$ and all
possible multiindices $J$ and
\begin{equation}
\label{2.4}
 \Phi^J_{\alpha} = D^J(\varphi_{\alpha}[u] - \xi^i [u] u^{\alpha}_i) +
\xi^i [u] u^{\alpha}_{J,i},
\end{equation}
compare this with formula (\ref{1.4}).\\

{\bf Note} there arise no convergence problems in defining the action
of an (infinitely) prolonged vector field on a function $F[u]$
since the latter only depends on a finite number of variables.\\

We now arrive at the definition of generalized symmetry. 
\end{df}

\begin{df}
A generalized vector field $V$ is a generalized symmetry of a system of
differential equations
\begin{displaymath}
 \Delta_j[u] = 0 \; \; (j=1,...,\l)
\end{displaymath}
if and only if
\begin{equation}
\label{2.5}
 pr(V)(\Delta_j) = 0 \; \; (j=1,...,\l)
\end{equation}
for solutions $u = f(x)$.\\

{\bf Note} it can be proved that for applications one has in mind
that condition (\ref{2.5}) results in
\begin{equation}
\label{2.6}
 pr(V)(\Delta_j) = \sum P_{k,j}^J [u]D^J (\Delta_k) \; \;
j=k=1,\ldots l \qquad |J|< \infty, \; \; P^J_{k,j}[u]\in C^\infty(J(x,u)) 
\end{equation}
or $pr(V)(\Delta_j) = 0$ when restricted to the manifold $Y \subset
J(x,u)$ defined by the system of differential equations and all its
differential consequences.\\

The concept of evolutionary or vertical vector field is a great
advantage in the computation of generalized symmetries.
\end{df}

\begin{df}
 A generalized vector field
\begin{equation}
\label{2.7}
 V = V_F = F_{\alpha}[u] \frac{\partial}{\partial u^\alpha}
\end{equation}
is called an evolutionary or vertical vector field.\\
The set of functions $(F_{\alpha})$ is called the {\em characteristic} of the
vector field $V$.\\

Note that for evolutionary vector fields we have a very elegant way for
the expression of the infinite prolongation (\ref{2.4}) i.e.
\begin{equation}
\label{2.8}
 pr(V_F) = D^J(F_{\alpha} [u]) \frac{\partial}{\partial u ^{\alpha} _ {J}}
\end{equation}
because $\xi^i [u] \equiv 0 \; \; (i=1,\ldots,p)$.\\
\end{df}

Moreover every infinitely prolonged vector field $V$ (\ref{2.2},\ref{2.3})
can be written as a sum of an evolutionary vector field and total
partial derivative vector fields i.e.
\begin{equation}
\label{2.9}
 pr(V) = pr(V_F) + \xi^i[u]D_i
\end{equation}
where the characteristic $F$ of the evolutionary vector field is given
by
\begin{equation}
\label{2.10}
 F_{\alpha}[u] = \varphi_{\alpha}[u] - u^{\alpha}_i \xi^i [u] \; \;
(\alpha=1,\ldots,q)
\end{equation}
Since vector fields $\xi^i[u]D_i$ satisfy the symmetry condition
(\ref{2.5}),(\ref{2.6}) in a trivial way; we can restrict the search for
generalized symmetries to the search for {\it evolutionary} vector fields.\\

To show the complexity of the computations involved in constructing
generalized symmetries we compute {\it third} order symmetries of the
potential form of Burgers' equation.

\begin{ex}
Burgers' equation is the following partial differential equation
\begin{equation}
\label{2.11}
 u_t = u_1^2 + u_2 \; \; (u_1=u_x,u_2=u_{xx})
\end{equation}
Note that differential consequences are given by
\begin{eqnarray}
\label{2.11b}
 u_{1t} & = & 2u_1u_2 + u_3 \qquad (u_{1t} = u_{xt})\\\nonumber
 u_{2t} & = & 2u_2^2 + 2u_1u_3 + u_4\\
 u_{3t} & = & 6u_2u_3 + 2u_1u_4 + u_5\nonumber
\end{eqnarray}
The characteristic of the evolutionary vector field $V_F$ is
\begin{equation}
\label{2.12}
 F[u] = F(x,t,u,u_1,u_2,u_3)
\end{equation}
Since we restrict to the solution manifold $u_t,u_{1t},..$ can be
eliminated by (\ref{2.11}),(\ref{2.11b})
\end{ex}
Now due to (\ref{2.8}) the symmetry condition (\ref{2.5}),(\ref{2.6})
reduces to
\begin{equation}
\label{2.13}
 D_tF - 2u_1 D_xF - D_x^2 F = 0
\end{equation}

\begin{equation}
\begin{array}{ll}
\label{2.14}
 \mbox{i.e. }& F_t + F_u(u_1^2 + u_2) + F_{u_1} (2u_1u_2 + u_3) +
  F_{u_2} (2u_2^2 + 2u_1u_3 + u_4)\\
 &+ F_{u_3} (6u_2u_3 + 2u_1u_4 + u_5)\\
 &- 2u_1(F_x + F_u u_1 + F_{u_1} u_2 + F_{u_2} u_3 + F_{u_3} u_4)\\
 &- \{F_{xx} + F_{xu} u_1 + F_{xu_1} u_2 + F_{xu_2} + F_{xu_3} u_4\\
 &+ u_1(F_{xu} + F_{uu} u_1 + F_{uu_1} u_2 + F_{uu_2} u_3 + F_{uu3} u_4)\\
 &+ u_2(F_{xu_1} + F_{uu_1} u_1 + F_{u_1u_1} u_2 + F_{u_1u_2}
   u_3 + F_{u_1u_3} u_4)\\
 &+ u_3(F_{xu_2} + F_{uu_2} u_1 + F_{u_1u_2} u_2 + F_{u_2u_2} u_3 +
   F_{u_2u_3} u_4)\\
 &+ u_4(F_{xu_3} + F_{uu_3} u_1 + F_{u_1u_3} u_2 + F_{u_2u_3} u_3 +
   F_{u_3u_3} u_4)\\
 &+ F_u u_2 + F_{u_1} u_3 + F_{u_2} u_4 + F_{u_3} u_5 \} = 0
\end{array}
\end{equation}

>From (\ref{2.14}) we see that the coefficient of $u_5$ vanishes
identically. The vanishing of the coefficients of $u_4,u_4^2$ lead to
\begin{eqnarray}
\label{2.15}
 u_4^2 : F_{u_3u_3} = 0
\end{eqnarray}
\begin{eqnarray}
\label{2.16}
 u_4 : -F_{xu_3} - u_1 F_{uu_3} - u_2 F_{u_1u_3} - u_3 F_{u_2u_3} = u_4
\end{eqnarray}

The first equation leads to the fact that $F_3$ is a polynomial of
degree $\leq 1$ in $u_3$ while the second equation results in

\begin{equation}
\label{2.17}
 F = \alpha(t)u_3 + \bar F(x,t,u,u_1,u_2)
 \end{equation}

Substitution of (\ref{2.17}) into (\ref{2.14}) leads to a polynomial
of degree 2 in $u_3$, the coefficients of which have to vanish i.e.
\begin{eqnarray}
\label{2.18}
 u_3^2 : \bar F_{u_2u_2} = 0
\end{eqnarray}
\begin{eqnarray}
\label{2.19}
 u_3 : \alpha'(t) + 6u_2\alpha(t) = 2 \bar F_{xu_2} + 2 \bar F{_uu_2} u_1 +
 2u_2 \bar F_{u_1u_2}
\end{eqnarray}
which results in

\begin{equation}
\label{2.20}
 \bar F(x,t,u,u_1,u_2) = 3\alpha u_1u_2 + (\frac{1}{2} \alpha'x +
\beta(t))u_2 + \tilde F(x,t,u,u_1)
\end{equation}
proceeding in this way we finally arrive at the fact that the solution
of (\ref{2.13}) is a linear combination of 10 vector fields whose
characteristics are given by

\begin{equation}
\label{2.21a}
 \begin{array}{rcl}
  F_0 &=& 1\\
  F_1 &=& u_1\\
  F_2 &=& tu_1 + \frac{1}{2} x\\
  F_3 &=& u_2 + u_1^2\\
  F_4 &=& t(u_2 + u_1^2) + \frac{1}{2} xu_1
 \end{array}
\end{equation}
\begin{equation}
\label{2.21b}
 \begin{array}{rcl}
  F_5 &=& t^2(u_2 + u_1^2) + txu_1 + (\frac{1}{2} t + \frac{1}{4} x^2)\\
  F_6 &=& u_3 + 3u_1u_2 + u_1^3\\
  F_7 &=& tF_6 + \frac{1}{2} x F_3\\
  F_8 &=& t^2F_6 + txF_3 + (\frac{1}{2} t + \frac{1}{4} x^2)F_1\\
  F_9 &=& t^3F_6 + \frac{3}{2} t^2 xF_3 + (\frac{3}{2} t^2 + \frac{3}{4}
  tx^2)F_1 + \frac{3}{4} tx + \frac{1}{8} x^3\\
 \end{array}
\end{equation}
  and\\
\begin{eqnarray*}
  F_{10} &=& \rho (x,t)e^{-u}
\end{eqnarray*}
whereas in (\ref{2.21b}) $\rho(x,t)$ is an arbitrary solution of the
heat equation $\rho_t = \rho_{xx}$.   

The existence of a symmetry (\ref{2.21b}) reflects the fact that the
equation at hand (\ref{2.11}) is in 1-1 correspondence with the heat
equation. The general theorem concerning this was proved by Kumei \&
Bluman \cite{KB}.\\
At the moment a number of computerprograms is available in
REDUCE,...,to handle the computations for symmetries p.e. \cite{K}.\\
In order to introduce the Lie bracket of generalized vector fields we
first prove the following lemma.

\begin{lemma}
 If $V_F$ is an evolutionary vector field then

\begin{equation}
\label{2.22}
 [pr(V_F),D_i] = 0
\end{equation}
interpreted as componentwise.
\end{lemma}

\noindent{\bf Proof.} First of all $\frac{\partial}{\partial
 u_j^\alpha} (D_iP) = \frac{\partial P}{\partial u_{J\backslash
 i}^\alpha} + D_i(\frac{\partial}{\partial u_J^\alpha}P)$\\
where $J\backslash i = (j_1,\ldots,j_{i-1},\ldots,j_p)$.\\
This implies that

\begin{equation}
\label{2.23}
 pr(V_F)(D_iP) = (D^J F_\alpha)\cdot \frac{\partial}{\partial
 u_J^\alpha}(D_iP) = (D^J F_\alpha)D_i(\frac{\partial}{\partial
 u_J^\alpha}P) + D^J F_\alpha \cdot \frac{\partial P}{\partial
 u^\alpha_{J\backslash i}} 
\end{equation}
We know that

\begin{equation}
\label{2.24}
 D_i(pr(V_F)P) = D_i((D^J F_\alpha) \cdot \frac{\partial P}{\partial
 u_J^\alpha}) = (D^J F\alpha)D_i(\frac{\partial
 P}{\partial u_J^\alpha}) + (D_iD_JF_\alpha) \frac{\partial P}{\partial
 u_J^\alpha}.   
\end{equation}

By changing summation index $J$ tot $J\backslash i$ we see that the
right hand sides in (\ref{2.23},\ref{2.24}) are equal, which proves
the Lemma.\\
As a corollary to this lemma we have
\begin{equation}
\label{2.25}
 pr(V_F)(D^JP) = D^J(pr(V_F)P) 
\end{equation}

\begin{th}
Let $V_Q,V_R$ be two evolutionary vector fields and $pr(V_Q)$,
$pr(V_R)$ their prolongations to $J(x,u)$ then the formal commutator is

\begin{equation}
\label{2.26}
 [pr(V_Q),pr(V_R)] = \tilde S
\end{equation}
where $\tilde S$ is the prolongation of an evolutionary vector field

\begin{equation}
\label{2.27}
 \tilde S = pr(V_S)
\end{equation}
and $S$ is defined by

\begin{equation}
\label{2.28}
 S_\alpha = pr(V_Q)(R_\alpha) - pr(V_R)(Q_\alpha) \qquad \alpha=1,\ldots,q
\end{equation}
\end{th}

\noindent{\bf Proof.} The definition of $S$ in (\ref{2.28}) is just
the computation of the $\frac{\partial}{\partial u^\alpha}$ component in
(\ref{2.26}).\\

The component of $\partial_{u_J^\alpha}$ in $\tilde S$ (\ref{2.26}) is
obtained from

\begin{displaymath}
 \tilde S_{u_J^\alpha} = pr(V_Q)D^J(R_\alpha) - pr(V_R)D^JQ_\alpha
\end{displaymath}
and by Lemma 2.1
\begin{displaymath}
 \tilde S_{u_J^\alpha} = D^J\{pr(V_Q)R_\alpha - pr(V_R)Q_\alpha\} =
D^J S_\alpha 
\end{displaymath}
stating that $\tilde S$ is just the prolongation of $V_S$ (cf.\ref{2.8}).\\
>From theorem 6 and the symmetry condition (3.5,6), we now have the
following

\begin{th}
the evolutionary generalized symmetries of a system of differential
equations
\begin{displaymath}
 \Delta_J[u]=0 \qquad (j=1,\ldots,\ell
\end{displaymath}
constitute a Lie algebra by the Lie bracket (\ref{2.26}).
\end{th}

\begin{ex} (Burgers' equations)
We compute some Lie brackets of evolutionary symmetries of example
3.4, (\ref{2.21a}),(\ref{2.21b}).\\
Take
\begin{eqnarray*}
 &Z_1=F_6=t(u_3+3u_1u_2+u^3_1) + \frac{1}{2}
          x(u_2+u^2_1)\\
 &X_1=u_1\\
 &X_2=u_2+u^2_1\\
 &X_3=u_3+3u_1u_2+u^3_1
\end{eqnarray*}
We now have the following result
\begin{eqnarray*}
 \left[V_{Z_1},V_{X_1}\right] &=& \frac{1}{2}V_{X_2}\\
 \left[V_{Z_1},V_{X_2}\right] &=& V_{X_3}\\
 \left[V_{Z_1},V_{X_3}\right] &=& \frac{3}{2}V_{X_4}
\end{eqnarray*}
where $V_{X_4}$ is a fourth-order generalized symmetry and
\begin{displaymath}
 X_4=u_4+3u^2_2+4u_1u_3+6u^2_1u_2+u^4_1
\end{displaymath}
In effect the generalized symmetries of example 3.4
(\ref{2.21a}),(\ref{2.21b}) constitute on {\em infinite dimensional
Lie algebra}.
\end{ex}

\setcounter{equation}{0}
\section{Nonlocal symmetries.}

Here we shall discuss special types of nonlocal symmetries as they
arise in certain special types of coverings. The notion of covering
has been introduced in \cite{KV} and \cite{KV2} (also called
Wahlquist-Estabrook prolongation) and has been discussed by
P. Gragert in his lecture \cite {G}.\\ 
For simplicity we restrict to two independent variables
$(x,t)(p=2)$.\\

In the discussion of coverings or prolongation one starts at the
infinite prolongation $Y$ of a $k$-th-order system of partial
differential equations, i.e. the original system together with all of
its differential consequences, defined on the infinite jet bundle
$J((x,t),u)$ i.e.

\begin{equation}
\label{3.1}
 D^J(\Delta_j[u]) = 0 \qquad j=1,\ldots,\ell,|J|<\infty
\end{equation}
 
An $s$-dimensional covering of (\ref{3.1}), with $(y_1,\ldots,y_s)$ as
local coordinates in the fibres, requires the existence of functions
\begin{displaymath}
 X_r([u],y_1,\ldots,y_s),T_r([u],y_1,\ldots,y_s) \qquad r=1,\ldots,s
\end{displaymath}
such that the extended or generalized total partial derivative
operators

\begin{eqnarray}
\label{3.2}
 \tilde D_x & = & D_x + X_r \frac{\partial}{\partial_{y_r}}\nonumber\\
            &   & \hspace{5cm} \mbox{(summation $r=1,\ldots,s$)}\\
 \tilde D_t & = & D_t + T_r \frac{\partial}{\partial_{y_r}}\nonumber
\end{eqnarray}
commute, i.e.

\begin{equation}
\label{3.3}
 [\tilde D_x,\tilde D_t] = 0
\end{equation}
which yields besides (\ref{3.1}) the covering condition

\begin{equation}
\label{3.4}
 \tilde D_xT_r - \tilde D_tX_r = 0 \; \; \; \mbox{ on (\ref{3.1}) }
\end{equation}
i.e.

\begin{equation}
\label{3.4a}
 D_xT - D_tX + [X,T] = 0
\end{equation}
where $X=(X_1,\ldots,X_s) T=(T_1,\ldots,T_s)$ and the bracket in
(\ref{3.4a}) is taken with respect to the fibre coordinates
$y=(y_1,\ldots,y_s)$.\\

As a special case we now consider coverings (\ref{3.2}),(\ref{3.4})
where $X_r,T_r$ are independent of $y=(y_1,\dots,y_s)$; (\ref{3.4})
then reduces to

\begin{equation}
\label{3.5}
 D_x(T_r) - D_t(X_r) = 0 \; \; \; \mbox{ on (\ref{3.1}) } (r=1,\ldots,s)
\end{equation}
i.e. $X_r,T_r$ determines a conservation law for (\ref{3.1})
and $Y_r=D_x^{-1}(X_r)$, as formal integral.\\

Analogously to (\ref{2.6}) we now introduce a {\em nonlocal} vertical
(generalized) vector field

\begin{equation}
\label{3.6}
 V_F = F_\alpha([u],y_1,\ldots,y_s) \frac{\partial}{\partial u_\alpha}
\end{equation}
and its prolongation to the infinite jetbundle

\begin{equation}
\label{3.7}
 pr(V_F) = \tilde D^J(F_\alpha([u],y_1,\ldots,y_s))
\frac{\partial}{\partial u_J^\alpha}
\end{equation}

We now define the notion of nonlocal symmetry.

\begin{df}

A nonlocal vector field $V_F$ (\ref{3.6}) determines a nonlocal
symmetry of (\ref{3.1}) if and only if

\begin{equation}
\label{3.8}
 pr(V_F)(\Delta_j) = 0 \; \; \; \mbox{ on (\ref{3.1}) },
j=1,\ldots,\ell
\end{equation}
where $pr(V_F)$ is defined by (\ref{3.7}).\\

{\bf Note:} The interested reader, comparing this definition with the
one given in Vinogradov \& Krasilshchik's work \cite{KV2}, might notice a
difference; in order to keep things simple and to outline the ideas we
just use this simplified definition.\\

We apply the notion of nonlocal symmetries to the construction of
nonlocal symmetries of the famous Korteweg-de Vries equation
(KdV-equation).
\end{df}

\begin{ex}
We start at the infinite prolongation of the KdV-equation i.e.,

\begin{equation}
\label{3.9}
 u_t = uu_1 + u_3 \qquad (u_1=u_x,u_3=u_{xxx})
\end{equation}
and its differential consequences.\\

If we apply the technique of the preceding section and search for
generalized symmetries of (\ref{3.9}) with characteristic
$F=F(u,u_1,\ldots,u_5)$, we arrive at the existence of

\begin{equation}
\label{3.10}
 \begin{array}{rclrcl}
 F_1 &=& u_1      &F_4 &=& 2u+xu_1+3t(uu_1+u_3)\nonumber\\
 F_2 &=& uu_1+u_3 &F_5 &=& 1+tu_1\\
 F_3 &=& \frac{5}{6} u_1u^2+\frac{10}{3}u_1u_2+\frac{5}{3}uu_3+u_5\nonumber
 \end{array}
\end{equation}
being the characteristics of $5$ generalized symmetries $V_{F_i} \;
(i=1,\ldots,5)$.\\
\end{ex}

Note that (\ref{2.3}),(\ref{2.4})

\begin{equation}
\label{3.11}
 \begin{array}{rcll}
 V_{F_1} &\doteq& \frac{\partial}{\partial x} , \qquad V_{F_2}
 \doteq \frac{\partial}{\partial t} &(x,t \mbox{-translation
 )}\\  
 V_{F_3} &\doteq& -x \frac{\partial}{\partial x} - 3t
 \frac{\partial}{\partial t} + 2u \frac{\partial}{\partial u}
 &(\mbox{scale transformation )}\\
 V_{F_4} &\doteq& t \frac{\partial}{\partial x} +
 \frac{\partial}{\partial u} &(\mbox{Gallilean Boost)}
\end{array}
\end{equation} 

It is an easy observation that $X_1=u, \; \; T_1=\frac{1}{2}u^2+u_2$
yield a conservation law for KdV-equation (\ref{3.9})

\begin{equation}
\label{3.12}
 D_xT_1-D_tX_1=0 \qquad \mbox{on (\ref{3.9}) }
\end{equation}

We introduce the {\em 1-dimensional covering} of (\ref{3.9}) with
$y=D_x^{-1}(u)$, and we are interested in the existence of a nonlocal
symmetry of (\ref{3.9}) i.e. solution of (\ref{3.8}) where

\begin{equation}
\label{3.13}
 H=H(x,t,u,\ldots,u_s,y)
\end{equation}
i.e.
\begin{equation}
\label{3.14}
 \tilde D_tH - u_1H - u \tilde D_xH - \tilde D_x^3H = 0
\end{equation}
\begin{eqnarray*}
 \tilde D_x &=& D_x + u\frac{\partial}{\partial y}\\
 \tilde D_t &=& D_t + (\frac{1}{2}u^2+u_2)\frac{\partial}{\partial y}
\end{eqnarray*}

Using an integration package the solution can be constructed in a
straightforward way, but since this would be very lengthy we proceed
in a more convenient way.\\
First of all, note that KdV-equation is graded due to the scale
transformation (\ref{3.11}).\\
i.e.
\begin{displaymath}
 [u]=2 \; ; \; [x]=-1 \; ; \; [t]=-3 \; ; \; [D_x]=1 \; ; \;
 [D_t]=3
\end{displaymath}
which implies

\begin{equation}
\label{3.15}
 [F_1]=3 \; ; \; [F_2]=5 \; ; \; [F_3]=7 \; ; \; [F_4]=2 \; ; \; [F_5]=0.
\end{equation}

We now search for a nonlocal symmetry whose characteristic is of
degree 4 and which is of polynomial degree 1 in $x,t,y$.\\

>From this we arrive at the Ansatz, based on the grading (\ref{3.15})

\begin{equation}
\label{3.16}
 H=t(F_3) + \alpha xF_2 + \beta {\bf y}u_1 + \gamma u_2 + \delta u^2u_1
\end{equation}
where $F_2,F_3$ are defined by (\ref{3.10}) and
$\alpha,\beta,\gamma,\delta$ constants to be determined. The symmetry
condition, due to the fact that $F_3,F_2$ satisfy (\ref{3.14})
themselves, reduces to

\begin{equation}
\label{3.17}
 \begin{array}{ll}
 &F_3 - \alpha uF_2 - 3\alpha D_x^2F_2 + \beta(\frac{1}{2}u^2+\beta
  {\bf y}(u_1^2+uu_2+u_4)\\
 &+ \gamma(3u_1u_2+uu_3+u_5) + 2\delta u(uu_1+u_3)-u_1(\beta
  {\bf y}u_1+\gamma u_2+\delta u^2)\\
 &-u(\beta uu_1+\beta {\bf y}u_2+\gamma u_3+2\delta uu_1)\\
 &-[4\beta u_1u_2+3\beta uu_3+\beta {\bf y}u_4+\gamma u_5+2\delta
  uu_3+6\delta u_1u_2] = 0
 \end{array}
\end{equation}

This condition leads to the following conditions for
$\alpha,\beta,\gamma,\delta$

\begin{equation}
\label{3.18}
 \begin{array}{rcl}
 u_5    &:& 1-3\alpha + \gamma-\gamma = 0\\
 uu_3   &:&
 \frac{5}{3}-\alpha-3\alpha+\gamma+2\delta-\gamma-3\beta-2\delta=0\\
 u_1u_2 &:&
 \frac{10}{3}-9\alpha+\beta+3\gamma-\gamma-4\beta-6\delta=0\\
 u^2u_1 &:&
 \frac{5}{6}-\alpha+\frac{1}{2}\beta+2\delta-\delta-\beta-2\delta=0 
 \end{array}
\end{equation}
or equivalently

\begin{equation}
\label{3.19}
 \begin{array}{l}
 1-3\alpha=0\\
 \frac{5}{3}-4\alpha-3\beta=0\\
 \frac{10}{3}-9\alpha-3\beta+2\gamma-6\delta=0\\
 \frac{5}{6}-\alpha-\frac{1}{2}\beta-\delta=0
 \end{array}
\end{equation}
solving (\ref{3.19}) we arrive at
\begin{displaymath}
 \alpha=\frac{1}{3} \; ,\; \beta=\frac{1}{9} \; ,\; \gamma=\frac{4}{3}
 \; ,\; \delta=\frac{4}{9}
\end{displaymath}
which leads to the characteristic ({\em nonlocal}) of a symmetry of
KdV-equation 

\begin{equation}
\label{3.20}
 H=tF_3+\frac{1}{3}xF_2+\frac{1}{9}yu_1+\frac{4}{3}u_2+\frac{4}{9}u^2u_1
\end{equation}

\setcounter{equation}{0}
\section{Recursion Operators and Nonlocal Symmetries.}

In this section we indicate the importance of nonlocal symmetries in
connection with the existence of recursion operators.\\
For simplicity we restrict to the case of (\ref{3.9}) two independent and one
dependent variable, keeping the KdV-equation as principal example in
mind.\\

Let us take a deeper look at the (generalized) symmetry condition
(\ref{2.5}),(\ref{3.8}) i.e.

\begin{equation}
\label{4.1}
 pr(V)(\Delta)=0 \mbox{ on } Y.
\end{equation}

If we use the prolongation formula (\ref{2.8}),(\ref{3.7}) it is a
straightforward procedure to see that the symmetry condition can be
rewritten as

\begin{equation}
\label{4.2}
 \sum(\frac{\partial\Delta}{\partial u_J}) D^J(F)=0
\end{equation}
which is reflected in (\ref{2.13}),(\ref{3.14}).\\
This observation urges us to introduce the socalled {\it linearization
operator} \cite{KV}

\begin{equation}
\label{4.3}
 \ell_\Delta = \sum(\frac{\partial\Delta}{\partial u_J}) D^J
\end{equation}
while (\ref{4.1}),(\ref{2.5}) can be written as

\begin{equation}
\label{4.4}
 \ell_\Delta F=0
\end{equation}

Suppose there exists a differential or integro- differential operator
$\cal R$, and associated to it some $\cal S$ such that the following relation
for operators, $\cal R,\cal S,$ hold

\begin{equation}
\label{4.5}
 \ell_\Delta {\cal R}={\cal S}\ell_\Delta
\end{equation}

Now assume that $F_0$ is a characteristic of a generalized symmetry of
$\Delta$ i.e.

\begin{equation}
\label{4.6}\
 \ell_\Delta F_0=0
\end{equation}

We then have

\begin{equation}
\label{4.7}
 \ell_\Delta({\cal R}F_0) = {\cal S}(\ell_\Delta F_0)=0
\end{equation}
i.e. ${\cal R}(F_0)$ is a characteristic of a generalized symmetry.\\

More generally, if an {\bf operator $\cal R$ satisfying (\ref{4.5})}
for some $\cal S$ exists then starting from a {\bf characteristic
$F_0$} of a symmetry we obtain a {\bf infinite hierarchy} (if not
zero) of generalized symmetrics whose {\bf characteristics} are
defined by 

\begin{equation}
\label{4.8}
 {\cal F}_n={\cal R}^n(F_0) \qquad n=0,\ldots
\end{equation}
Such an operator $\cal R$ is called a {\bf recursion operator} for
generalized symmetries

\begin{ex}
The KdV-equation

\begin{equation}
\label{4.9}
 \Delta(u)=u_t-uu_1-u_3=0
\end{equation}
admits a recursion operator for symmetries

\begin{equation}
\label{4.10}
 {\cal R}=D_x^2 + \frac{2}{3}u + \frac{1}{3}u_1D_x^{-1}
\end{equation}
where $D^{-1}_x$ has to be understood as a formal integral [1].
It is a somewhat tedious calculation to show that

\begin{equation}
\label{4.11}
 \ell_\Delta {\cal R} = {\cal R}\ell_\Delta
\end{equation}
i.e. ${\cal S}={\cal R}$.\\

If we start with $F_1=u_1$ then

\begin{equation}
\label{4.12}
\begin{array}{rcccccl}
 F_2 &=& {\cal R}F_1   &=& {\cal R}u_1 &=& uu_1 + u_3\\
 F_3 &=& {\cal R}^2F_1 &=& {\cal R}F_2 &=&
 \frac{5}{6}u_1u^2+\frac{10}{3}u_1u_2+\frac{5}{3}uu_3+u_5
\end{array}
\end{equation}
and so on.
\end{ex}

We are now in a position underlign the importance of the notion of
nonlocal symmetry.\\
First of all, if we would apply the recursion formula (\ref{4.8})
starting at $F_4$ or $F_5$ (in effect $F_4={\cal R}F_5$) then in order to
compute ${\cal R}F_4$ we would have to allow nonlocal variables $y$ to come
in.\\
Moreover the nonlocal characteristic $H$ (\ref{3.20}) is just nothing
else but

\begin{equation}
\label{4.13}
 H=3{\cal R}F_4
\end{equation}

Secondly, if we compute the generalized Lie-Bracket for generalized
vector field (\ref{2.28}) and compute Lie-Brackets with the non local
vector field $V_H$ we arrive at

\begin{equation}
\label{4.14}
 \begin{array}{rcl}
 \left[V_H,V_{F_1}\right] &=& c_1F_2\\
 \left[V_H,V_{F_2}\right] &=& c_2F_3 
 \end{array}
\end{equation}
$c_1,c_2$ being some nonzero constants.\\

In effect the nonlocal generalized symmetry $V_H$ acts as {\bf
recursion symmetry}.\\

{\bf Final Remarks.}\\
In this lecture I have tried to give you an introduction to and an
impression of the beautiful world of symmetries of differential
equations, where a lot of research is needed to explore the beautiful
structures in this field of applied mathematics.\\

For the interested reader I would recommend the book of Bluman-Kumei
\cite{KB} as a starting point, the book by Olver as a rigorous and
deep discussion of all the mathematics involved, and the work of my
Russian friends Vinagradov, Krasil'shchik for the beautiful and rich
geometrical structures underlying all the notions.

\begin{thebibliography}{999}
\bibitem{O} Olver P.J.A., Applications of Lie Groups to Differential
Equations Graduatem Texts in Mathematics 107. Springer Verlag, New
York-Berlin-Heidelberg (1986).
\bibitem{V} Vinagradov A.M., Local symmetries and conversation laws.
Acta Applicandae Mathematicae Vol 3 (1984), pp. 21-78.
\bibitem{KV} Krasil'shchik I.S. \& Vinagradov A.M., Nonlocal
symmetries and the theory of coverings. Acta Applicandae Mathematicae
Vol 3 (1984), pp. 79-96.
\bibitem{KV2} Krasil'shchik I.S. \& Vinagradov A.M., Nonlocal Trends
in the Geometry of Differential Equations: Symmetries, Conservation
Laws, and B\"{a}cklund Transformations Acta Applicandae Mathematicae
Vol 15 (1989), pp. 161-209.
\bibitem{K} Kersten P.H.M., Infinitesimal Symmetries: a Computational
Approach C.W.I. Tract 34. Centre for Mathematics and Computer Science,
Amsterdam (1987).
\bibitem{G} Gragert P.K.H., Prolongation algebras of nonlinear PDE,
These Notes. 
\bibitem{KB} Kumei S. \& Bluman G., Symmetries and Differential
Equations. Applied Mathematical Sciences 81. Springer Verlag, New
York-Berlin-Heidelberg (1989). 
\end{thebibliography}

\end{document}