-
Notifications
You must be signed in to change notification settings - Fork 1
/
binary.tex
1101 lines (953 loc) · 49 KB
/
binary.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
\documentclass{article}
\usepackage[utf8]{inputenc}
\usepackage[T1]{fontenc}
\usepackage{unicode}
\usepackage{math}
\usepackage[margin=25mm]{geometry}
\usepackage{color}
\usepackage[all]{xypic}
\DeclareUnicodeCharacter{26A0}{{\color{red}\ensuremath{\lower .25ex\hbox{\Large
$\triangle$\hskip -1.25ex}!\;\,}}}
\def\transpose#1{{\vphantom{#1}}^{\mathrm{t}}\!#1}
\def\mat#1{\begin{pmatrix}#1\end{pmatrix}}
\def\F{\mathbb{F}}
\def\appli#1#2#3#4{\begin{array}{rcl}#1&\longrightarrow\\
#3&\longmapsto\end{array}}
\def\operp{\mathrel{\vbox{\offinterlineskip\ialign{\hfil##\hfil\cr
$\scriptstyle\perp$\cr\noalign{\kern .3ex}$\scriptstyle\bigoplus$\cr}}}}
\DeclareMathOperator\GL{GL}
\DeclareMathOperator\Quad{Quad}
\DeclareMathOperator\Bil{Bil}
% \DeclareMathOperator\Sym{Sym}
\DeclareMathOperator\Alt{Alt}
\def\kR{_{(k,R)}}
% \def\chk#1{#1^{\scriptscriptstyle\vee}}
\usepackage{graphicx}
\def\chk#1{#1^{\smash{\scalebox{.7}[1.4]{\rotatebox{90}{\guilsinglleft}}}}}
\let\fr\mathfrak
\long\def\note#1{{\color{magenta}{\textbf{Note:} #1}}}
\begin{document}
\title{Solving IP1S in characteristic two: the fully irregular case}
A pencil is \emph{fully irregular} if its regular part is zero,
\emph{i.e.} if it is isomorphic to a direct sum of Kronecker modules.
\section{The equations for IP1S}
Let~$k$ be a finite field with characteristic~$2$. Except where
indicated, all tensor products shall be taken over~$k$.
\subsection{Kronecker modules as homogeneous polynomials}
For~$d ≥ 0$, we write~$H_d$ for the $d+1$-dimensional space of
homogeneous polynomials in~$k[x:y]$ of degree~$d$. For any~$h ∈ H_m$, the
multiplication by~$m$ defines linear maps~$H_d → H_{d+1}$, which we again
write~$h$; all these maps
commute with each other. We write~$\chk{H_d}$ for the dual vector space
of~$H_d$, and~$\chk{h}: \chk{H_{d+1}} → \chk{H_{d}}$ for the transposed
maps of~$h$. We also write~$\chev{φ, f}$ for the canonical bilinear
map~$H_d ⊗ \chk{H_d} → k$; we note that, for~$f ∈ H_{d-m}$ and~$φ ∈
\chk{H_d}$, we always have the relation~$\chev{φ, fh} = \chev{\chk{h} φ, f}$.
The \emph{Kronecker module} of degree~$d ≥ 0$ over~$k$ is the
$2d+1$-dimensional space~$K_d = H_{d-1} ⊕ \chk{H_d}$, equipped with the
symmetric bilinear pencil~$(b_{t})$ defined for~$f, f' ∈ H_{d-1}$ and~$φ,
φ' ∈ \chk{H_{d}}$ by
\begin{equation}
b_t (f, f') \;=\; b_t (φ, φ') \;=\; 0; \qquad
b_t (f, φ) \;=\; \chev {φ, (ty-x) f}.
\end{equation}
This means that $b_{0} (f, φ) = \chev {φ, xf} = \chev {\chk{x} φ, f}$,
and $b_{∞} (f, φ) = \chev{φ, yf} = \chev{\chk{y} φ, f}$. Since $f ∈
H_{d-1}$ and~$φ ∈ \chk{H_{d}}$, we have~$xf ∈ H_{d}$ and~$\chk{x} φ ∈
\chk{H_{d-1}}$, so these pairings are well-defined.
% For~$d ≥ -1$, we write~$P_d$ for the
% $d+1$-dimensional space of polynomials in~$k[x]$ of degree~$≤ d$,
% and~$P_d^∨$ for its dual. Let~$ι_d: P_{d-1} → P_{d}$ be the natural
% injection map and~$x_d: P_{d-1} → P_{d}$ be the multiplication by~$x$.
% The various maps~$ι_d$ and~$x_d$, when composable, commute with each
% other, and we shall simply write them as~$ι$ and~$x$. They define
% transposed maps~$\chk{ι}$ and~$\chk{x}: P_{
% The \emph{Kronecker module} of degree~$d$
% over~$k$ is then the space~$K_d = P_{d-1} ⊕ \chk{P_d}$, equipped with the
% symmetric bilinear pencil~$(b_{t})$ defined for~$f, f' ∈ P_{d-1}$ and~$φ, φ' ∈
% \chk{P_{d}}$ by
% \begin{equation}
% b_t (f, f') \;=\; b_t (φ, φ') \;=\; 0, \qquad
% b_t (f, φ') \;=\; \chev{φ, (x - t) ι_d f}.
% \end{equation}
\subsection{Kronecker modules with coefficients}
A \emph{coefficient space} is a bilinear space~$E$ isomorphic to~$k^n$,
together with its standard scalar product~$u, v ↦ u · v$. For any linear
maps~$α: V → E$, $β: W → E$, we write~$α · β$ for the corresponding
bilinear form on~$V × W$.
\note{En première lecture, on pourra ignorer ces espaces de coefficients,
c'est-à-dire supposer que tous les coefficients~$α$, $β$ sont scalaires.}
The bilinear module~$K_d^{n}$ is isomorphic to~$E ⊗ K_d$. For~$u ∈ E ⊗
H_m$ written as~$u = ∑ u_i ⊗ x^i$ with~$u_i ∈ k^n$ and~$x^i ∈ H_m$, we
define, for~$f ∈ H_d$ and~$φ ∈ \chk{H_d}$,
\begin{equation}
u f \;=\; ∑ u_i ⊗ (x^i f) \;∈ E ⊗ H_{d+m}, \quad
\chk{u} φ \;=\; ∑ u_i ⊗ (\chk{(x^i)} φ) \;∈ E ⊗ \chk{H_{d-m}}.
\end{equation}
Likewise, for any~$h ∈ \Hom(E, E') ⊗ H_m$, write~$h = ∑ h_i ⊗ x^i$
with~$h_i ∈ \Hom (E, E')$. For any~$u ⊗ f ∈ E ⊗ H_d$, we define~$h(u ⊗
f) ∈ E' ⊗ H_{d+m}$ by~$h(u ⊗ f) = ∑ h_i(u) ⊗ (x^i f)$. An element
of~$\Hom (E ⊗ H_d, E' ⊗ H_{d+m})$ defined in this way is called
\emph{principal}. In the same way, we define principal elements of~$\Hom
(E ⊗ \chk{H_{d+m}}, E' ⊗ \chk{H_{d}})$.
Let~$r ≤ s$ be two integers and~$E, E'$ be two coefficient spaces. We say
that a bilinear form~$B$ on~$(E ⊗ H_r) × (E' ⊗ \chk{H_s})$ is
\emph{principal} if it is of the form~$B(f, φ) = \chev {φ, hf} = \chev
{\chk{h} φ, f}$ for some~$h ∈ \Hom (E, E') ⊗ H_{s-r}$.
\begin{lem}\label{lem:principal}
Let~$r ≤ s$ be two integers and~$E, E'$ be two coefficient spaces.
\begin{enumerate}
\item A bilinear form~$B$ on~$(E ⊗ H_r) × (E' ⊗ \chk{H_s})$ is principal
if, and only if, for any polynomial~$z$, $B(zf, φ) = B(f, \chk{z} φ)$.
\item Let~$α: E ⊗ H_r → E' ⊗ H_s$ and~$β: E' ⊗ \chk{H_s} → E ⊗ \chk{H_r}$ be
two linear maps. If $1 · α + β · 1$~is principal, then both~$α$ and~$β$
are principal.
\end{enumerate}
\end{lem}
\subsection{The automorphisms of a direct sum of Kronecker modules}
Any totally irregular pencil of quadrics is isomorphic to a space of the
form~$⨁ E_d ⊗ K_d$, where~$E_d = k^{n_d}$, and $(n_d)$~is some finite
sequence of integers.
\begin{prop}\label{prop:aut-Ed-Kd}
The automorphisms of~$⨁ E_d ⊗ K_d$ are exactly the maps of the form
\[ u ⊗ f ↦ ∑_{d' ≥ d} α_{d',d} (u) f + ∑_{d'} \chk{f} γ_{d',d} (u), \quad
u ⊗ φ ↦ ∑_{d' ≤ d} \chk{β_{d',d} (u)} φ, \]
where~$α_{d',d} ∈ \Hom (E_d, E_{d'}) ⊗ H_{d'-d}$, $β_{d',d} ∈ \Hom (E_d,
E_{d'} ⊗ H_{d-d'})$ and~$γ_{d',d} ∈ \Hom (E_{d}, E_{d'}) ⊗
\chk{H_{d+d'-1}}$ satisfy the following relations: let~$A, B, C$ be the
matrices~$(α_{d',d})$, $(β_{d',d})$, and~$(γ_{d',d})$; then
$\transpose{B} · A = 1$, and $\transpose{C} · A$~is antisymmetric.
\end{prop}
% \begin{prop}\label{prop:aut-Ed-Kd}
% The group of automorphisms of~$⨁ E_d ⊗ K_d$ is generated by the following
% elements:
% \begin{enumerate}
% \item elements of the form~
% \begin{equation}
% u ⊗ f ↦ ∑_{d' ≥ d} α_{d, d'} (u) f, \quad
% u ⊗ φ ↦ ∑_{d' ≤ d} \chk{β_{d, d'} (u)} φ,
% \end{equation}
% where~$α_{d, d'}: E_{d} → E_{d'} ⊗ H_{d'-d}$ and~$β_{d,d'}: E_{d} → E_{d'} ⊗ H_{d-d'}$ are linear maps satisfying
% the relations, for~$d ≤ d'$
% \begin{equation}
% ∑_{i=d}^{d'} α_{d,i} · β_{d',i} =
% \begin{cases} 1 & \text{if $d = d'$,}\\0 & \text{else.}\end{cases}
% \end{equation}
% \item elements of the form~$u ⊗ φ ↦ u ⊗ φ$, $u ⊗ f ↦ ∑_{d'} γ_{d,d'} (u)
% f$, where~$γ_{d,d'}: E_d → E_{d'} ⊗ \chk{\P_{d+d'}}$ is a linear map, such
% that~$γ_{d,d'} = \transpose{γ_{d', d}}$.
% \item elements of the form
% \begin{equation}
% u ⊗ φ ↦ u ⊗ φ, u ⊗ f ↦ u ⊗ f + ∑_{d'} \chk{f} γ_{d,d'} (u),
% \end{equation}
% where~$γ_{d,d'}: E_{d} → E_{d'} ⊗ \chk{H_{d+d'-1}}$ is such
% that~$\transpose{γ_{d,d'}} = γ_{d',d}$.
% \end{enumerate}
% \end{prop}
In the above theorem, we understand the elements~$u ⊗ f$ and~$u ⊗ φ$ to
belong to the spaces~$E_d ⊗ H_{d-1}$ and~$E_d ⊗ \chk{H_d}$. For~$d > d'$,
the space~$H_{d' - d}$ is zero, and therefore~$α_{d,d'} = β_{d',d} = 0$.
The relation~$\transpose{A} · B = 1$ means that, for all~$d, d'$, $∑_i
\transpose{α_{i,d}} · β_{i,d'} ∈ \chk{E_d} ⊗ \chk{E_{d'}} ⊗ H_{d'-d}$ is the standard
scalar product of~$E_d$ if $d = d'$ and~$0$ else; this product is to be
interpreted as the collection of the corresponding terms of all degrees
in~$(x:y)$, and the scalar product is taken in~$E_i$. Therefore, the
elements~$α_{i,j}$ uniquely determine all of the~$β_{i,j}$. Likewise, the
relation~$\transpose{A} · C + \transpose{C} · A = 0$ means that $∑_i
\transpose{α_{i,d}} · γ_{i,d'} + \transpose{γ_{i,d}} · α_{i,d'} = 0 ∈ E_d
⊗ E_{d'} ⊗ H_{d+d'-1}$, where scalar products are taken in~$E_i$.
\begin{proof}[(Sketch of) Proof of Prop.~\ref{prop:aut-Ed-Kd}]
Let~$F$ be an automorphism of~$V = ⨁ E_d ⊗ K_d$. The space~$V$ is
filtered by the degrees of its kernel vectors: for all~$d$, the
sub-space~$⨁ E_{d'} ⊗ \chk{H_{d'}}$ for~$d' ≤ d$ is exactly the linear span
of all kernel vectors of degree~$≤ d$. Therefore, $F$~is upper triangular
on these spaces, so that we may write, for~$u ∈ E_d$, $φ ∈ \chk{H_{d}}$
and~$f ∈ H_{d-1}$:
\begin{equation}
F(u ⊗ f) = ∑_{d'} α_{d',d} (u ⊗ f) + ∑_{d'} γ_{d',d} (u ⊗ f), \quad
F(u ⊗ φ) = ∑_{d' ≤ d} β_{d',d} (u ⊗ φ),
\end{equation}
where~$α_{d',d}: E_d ⊗ H_{d-1} → E_{d'} ⊗ H_{d'-1}$, $γ_{d',d}: E_d ⊗
H_{d-1} → E_{d'} ⊗ \chk{H_{d'}}$ and~$β_{d',d}: E_d ⊗ \chk{H_{d}} → E_d ⊗
\chk{H_{d'}}$. Moreover, for all~$d$, since $β_{d,d}(u ⊗ φ)$~preserves the
kernel of~$b$, it is principal, i.e. of the form~$β_{d,d,0} (u) ⊗ φ$
where $β_{d,d,0}: E_d → E_d$~is a linear map.
The relation~$b(F (u ⊗ f), F(v ⊗ φ)) = b(u ⊗ f, v ⊗ φ)$ implies that the
matrix~$A$ of all endomorphisms~$α$ is the inverse transpose of the
matrix~$B$ of the endomorphisms~$β$; since $B$~is upper triangular,
$A$~is lower triangular. We may then use Lemma~\ref{lem:principal} to
prove that all the~$α_{d',d}$ and~$β_{d',d}$ are principal, by induction
on~$d-d'$.
% Write~$β · α$ for the bilinear product of~$β(v ⊗ φ)$ and~$α(u ⊗ f)$, so
% that we have for all~$r, s$ the relation $∑_{i ≤ s} β_{s,i} · α_{r,i}
% = 1$ if $r = s$ and $0$~else.
% We now prove by induction on~$d$ the following statement: $β_{s,s} ·
% α_{s,s} = 1$ and, for all~$r > s$, we have~$α_{r,s} = 0$. For~$s = 0$,
% the previous relation becomes~$β_{0,0} α_{r,0} = 1$ if~$r = 0$ and
% $0$~else. The case~$r = 0$ proves that~$β_{0,0}$~is invertible and
% therefore that all~$α_{r,0}$ for~$r ≥ 1$ are zero. The induction step
% works in the same way: the relation for~$s$ is now~$∑
Finally, the relation on the coefficients~$γ_{d',d}$ comes directly from
the relation $b(F (u ⊗f), F(u' ⊗ f')) = 0$.
\end{proof}
\subsection{Action on the diagonal coefficients}
Let~$(V, b) = ⨁ E_d ⊗ K_d$ be a totally irregular bilinear pencil and
let~$q$ be a quadratic pencil on~$V$ whose polar form is~$b$. Since $b$~is
isotropic on~$⨁ E_d ⊗ H_{d-1}$ and on~$⨁ E_d ⊗ \chk{H_d}$, for any
element~$z = ∑ u_i ⊗ f_i + ∑ v_j ⊗ φ_j ∈ V$, we have
\begin{equation}
q(z) = ∑_i q(u_i ⊗ f_i) + ∑_j q(v_j ⊗ φ_j) +
∑_{i,j} b(u_i ⊗ f_i, v_j ⊗ φ_j).
\end{equation}
Let~$σ$ be the absolute Frobenius defined by~$σ(x)=x^2$. We note that the
restriction of~$q$ to one of the spaces~$E_d ⊗ H_{d-1}$ or~$E_d ⊗
\chk{H_{d}}$ is a $σ$-linear form. The quadratic form~$q$ is determined by the
restrictions of~$σ^{-1} ∘ q$ to these spaces, which are linear forms.
For~$u ∈ E_d$, $f ∈ H_{d-1}$ and~$φ ∈ \chk{H_d}$, we write $λ_d(u ⊗ f) =
σ^{-1} ∘ q\: (u ⊗ f)$ and~$μ_d (u ⊗ φ) = σ^{-1} ∘ q\: (u ⊗ φ)$.
% which we write~$λ_d ∈ E_d ⊗
% \chk{H_{d-1}}$ and~$μ_d ∈ E_d ⊗ H_d$ (using the scalar product to
% identify~$E_d$ with its dual).
Let~$F$ be an automorphism of~$b$, and define its coefficients
$α_{d,d'}$, $β_{d,d'}$ and~$γ_{d,d'}$ as in Prop.\ref{prop:aut-Ed-Kd}
above.
The action of~$F$ on the quadratic pencil~$q$ is given as follows:
\begin{align*}
q ∘ F\;(u ⊗ f) &= ∑_i q(α_{i,d} (u) f) +
∑_i q(\chk{f} γ_{i,d} (u)) + ∑_i b(γ_{i,d} (u), α_{i,d} (u) f^2);\\
q ∘ F\;(u ⊗ φ) &= ∑_i q(β_{i,d} (u) φ).
\end{align*}
Define as above~$λ_d$, $μ_d$ as the diagonal parts of~$q$. Also
let~$γ'_{d,d'} = ∑_i \transpose{α_{i,d}} γ_{i,d'} ∈ \Hom (E_d, E_{d'}) ⊗
\chk{H_{d+d'-1}}$, so that the anti-symmetry condition on the
coefficients~$γ$ becomes $\transpose{γ'_{d,d'}} = γ_{d',d}$.
Write~$γ'_{d,d',j} = ∑_{i,r} \transpose{α_{i,d,r}} · γ_{i,d',r+j}$ for
the coefficients of~$γ'$, as a linear map on~$H_{d+d'-1}$.
The diagonal coefficients~$λ'_d$, $μ'_d$ of~$q' = q ∘ F$ are given by
\begin{align*}
λ'_d &= ∑_i λ_i ∘ α_{i,d} + ∑_i μ_i ∘ γ_{i,d}
+ ∑_{i,j} σ^{-1} (γ'_{d,d,2j+1} - t γ'_{d,d,2j}) ξ_{j,d-j},\\
μ'_d &= ∑_i μ_i ∘ β_{i,d}.
\end{align*}
\begin{prop}\label{prop:ip1s-semilinear}
The $n$-dimensional IP1S problem for totally irregular pencils over a
field of characteristic two is equivalent to linears equation of
dimension~$O(n^2)$ and one semi-linear equation of
the form
\begin{equation*}
X = A σ(X) + B,
\end{equation*}
where $A$~is a square matrix and $X, B$~are column matrices of
dimension~$O(n^2)$.
\end{prop}
\section{Solving Frobenius equations}
% Part~1 shows how to bring the IP1S problem to an equation of the form
% \begin{equation}\label{eq:semi-linear}
% X = A σ(X) + B,
% \end{equation}
As the matrix~$A$ of Proposition~\ref{prop:ip1s-semilinear} seems to be
quite general, we give a generic method for this family of
\emph{Frobenius equations}. In the IP1S problem, the base field~$k$ has
characteristic~two; we present here the general case for any base field.
Let~$k[φ]$ be the non-commutative ring of polynomials in~$φ$, with the
relations~$φ c = σ(c) φ$ for all~$c ∈ k$. This ring is Euclidean.
More precisely, the (left-side) Euclidean algorithm works: given two
elements~$a, b ∈ k[φ]$, there exist elements~$u, v$ such that~$u a + v b =
d$ where $d$~is the gcd of~$a$ and~$b$.
Moreover, the only two-sided ideals of~$k[φ]$ are those generated by the
powers of~$φ$.
From these two remarks and [Jacobson, Th.~19] we get the following.
\begin{thm}
Let~$φ: k^n → k^n$ be a semi-linear endomorphism. There exists a basis
of~$k^n$ in which the matrix of~$φ$ is the direct sum of cyclic matrices,
at most one of which is not nilpotent.
\end{thm}
Assume now that $A$~is cyclic and let~$R = k[a]/f(a)$ be the $k$-algebra
generated by~$a$; the equation~\eqref{eq:semi-linear} may then be written
as
\begin{equation}\label{eq:semi-linear-pol}
x = a σ(x) + b,
\end{equation}
where~$a$, $b$ and~$x$ belong to the finite algebra~$k[a]$, and $σ$~is
the absolute Frobenius automorphism. To the primary factorization~$f = ∏
f_i$ of~$f$, there corresponds a factorization~$R = ∏ R_i$ where $R_i$~is
a local algebra. Using Chinese remainders, we may therefore assume that
$R$~is a local algebra and~$f = f_0^d$, where $f_0$~is irreducible.
We first see to the case where~$d = 1$, \emph{i.e.} $R$~is an extension
of the field~$k$. Write~$k_0 = \F_p$ be the field fixed by the
Frobenius~$σ$ and let~$N_0 = N_{R/k_0}$ and~$\Tr_0 = \Tr_{R/k_0}$ be the
norm and trace operators. The equation~\eqref{eq:semi-linear-pol} implies
that
\begin{equation}
x = b' + N_0(a) x, \quad\text{where $b' = b + a σ(b) + … + a σ(a) …
σ^{n-2}(a) σ^{n-1}(b)$.}
\end{equation}
If $N_0(a) ≠ 1$, then this gives as a unique solution~$x =
b'/(1-N_0(a))$. If, on the contrary, $N_0(a) = 1$, then by
Hilbert's theorem~90~[Lang, VI.6.1], there exists~$u ∈ R^{×}$ such
that~$a = σ(u)/u$; then~$x' = ux$ satisfies the equation~$x' = σ(x') +
ub$. This last equation, using the additive form of Hilbert's theorem~90
[Lang, VI.6.3], has a solution if, and only if, $\Tr_0 (ub) = 0$. We
note that both forms of Hilbert's theorem are algorithmic.
In the general case where~$d ≥ 1$, let~$\fr m$ be the maximal ideal
of~$R$. We may use the preceding paragraph to compute a solution~$x_0$ of
the equation modulo~$\fr m$. Moreover, we notice that~$f(x) = x - a σ(x)
- b$ is a polynomial and that~$f'(x) = 1$; therefore, this polynomial is
separated, and we may use Hensel's lemma to lift the approximate
solution~$x_0$ to a full solution.
\section{An intrinsic version of regular pencils}
This is an attempt to give a more intrinsic presentation of §2
of the IP2S document
\texttt{isomorphism\_polynomials}.
\subsection{Bilinear pencils}
Let~$b_{λ} = b_{∞} (λ - c)$ be a regular (principal) bilinear pencil
on~$V$. Using primary factorization, we assume that the characteristic
polynomial~$f$ of~$c$ is primary, \emph{i.e.} of the form~$f = p^d$
with~$p$ irreducible. We define the local algebras~$R_n = k[c]/p(c)^n$
and the extension field~$K = R_1 = k[c]/p(c)$. For each~$n$, $R_n$~is
isomorphic to~$K[π]/π^n$.
\note{pour prouver l'isomorphisme, il suffit de construire un
morphisme~$K → R_n$, càd de résoudre l'équation~$p(x) = 0$ dans~$R_n$.
Puisque $c$~est une solution approchée, et $p$~étant séparable, le lemme
de Hensel prouve qu'il existe une solution exacte.}
The form~$b_{∞}$ is $k$-bilinear on~$V$, and commutes with~$R_n$ in the
following sense: for all~$a ∈ R_n$, we have~$b_{∞} (x, ay) = b_{∞} (ax,
y)$. In particular, it commutes with~$K$. Moreover, morphisms in the IP1S
sense correspond to $R$-linear maps. We therefore define the category of
$(k, R)$-bilinear modules to be the category whose objects are
$R$-modules of finite type equipped with bilinear forms commuting
with~$R$, and whose maps are $R$-linear maps.
% We write~$\Bil_R (M)$, $\Sym_R (M)$ and $\Quad_R (M)$ for the $R$-modules
% of bilinear, symmetric and quadratic forms, and $\Bil\kR (M)$, $\Sym\kR
% (M)$ and $\Quad\kR (M)$ for the $k$-vector spaces of forms commuting
% with~$R$. We say that a quadratic form commutes with~$R$ if its polar
% form does.
\begin{prop}\label{prop:trace-form}
Let~$K/k$ be a separable field extension, and~$V$ be a finite-dimensional
vector space over~$K$.
For any $k$-bilinear form~$b: V ⊗ V → k$ commuting with~$K$, there
exists a unique $K$-bilinear form~$b_K: V ⊗ V → K$ such that~$b =
\Tr_{K/k} \,∘ \,b_K$.
\end{prop}
\begin{proof}
We first recall the (classical) proof of the result when~$V = K$. In this
case, let~$z$ be a primitive element of~$K$ and write $d = [K:k]$.
Since~$K/k$ is
separable, the trace form is non-degenerate~\cite[VI~5.2]{lang} and there
exists a unique element~$c ∈ K$ such that $\Tr (c z^i) = b(1, z^i)$ for
all~$i=0, …, d-1$. We immediately see that, for all~$x, y ∈ k$, $\Tr (c x
y) = b(x, y)$.
The general case now follows directly from taking coordinates in~$K$: all
coordinates of a bilinear form are themselves bilinear forms.
% We now prove the general case. Let~$n = \dim_K V$ and write the matrix
% of~$b$ as~$(b_{i,j})$, where $b_{i,j}: K ⊗ K → k$ is $k$-bilinear. Since
% $b_{i,j}$ again satisfies the hypotheses of the proposition, there exists
% a unique~$c_{i,j} ∈ K$ such that~$b_{i,j} (x,y) = \Tr_{K/k} (c_{i,j} x
% y)$. The bilinear form~$b_K$ is then the form with matrix~$(c_{i,j})$.
\end{proof}
\note{Noter que la
construction de Gilles pour la matrice~$T$ telle que $T M = \transpose{M}
T$ revient à peu près à la construction de la forme trace~$x,y ↦
\Tr_{K/k} (xy)$, et est donc aussi contenue dans ce résultat.}
\note{il n'est peut-être pas nécessaire que $p$~soit irréductible, mais
seulement séparable, de même que pour Hensel ci-dessus. On pourrait donc
vraisemblablement travailler directement avec les diviseurs
élémentaires ; mais je suis plus à l'aise avec des corps finis que des
algèbres séparables...}
For any integer~$m$, we define~$R_m$ as the local $k$-algebra $R_m =
k[π]/π^m$, and the $k$-linear form~$τ_m$ on~$R_m$ as the coefficient
of~$π^{d-1}$. We write~$τ$ instead of~$τ_m$ where there is no ambiguity.
We note that~$τ(xy)$, as a $k$-bilinear form on~$R_m$, is non-degenerate.
Actually, multiplication by~$π^{d-1}$ defines a (non-canonical)
$R$-linear isomorphism between~$R$ and its dual (as a $k$-vector
space)~$R^{∨}$.
\begin{prop}\label{prop:trace-local}
Let~$R = R_m$ and $M, N$ be $R$-modules of finite length. For any
$k$-bilinear form~$b: M ⊗ N → k$ commuting with~$R$, there exists a
unique $R$-bilinear form~$b_R: M ⊗ N → R$ such that~$b = τ ∘ b_R$.
% Let~$R$ be the local algebra~$R = k[π]/π^d$ and define the $k$-linear
% form~$τ_d$ on~$R$ as the coefficient of~$π^{d-1}$. For any finite free
% $R$-modules~$V, W$ and $k$-bilinear form~$b: V ⊗ W → k$ commuting
% with~$R$, there exists a unique $R$-bilinear form~$b_R: V ⊗ W →
% R$ such that~$b = τ_d ∘ b_R$.
\end{prop}
Note that we do not demand that $M$, $N$ be \emph{free} as $R$-modules.
Moreover, the unicity of~$b_R$ shows that both forms~$b$ and~$b_R$ are
simultaneously symmetric or antisymmetric. However, when $k$~is a field
of characteristic two, it may happen, as we shall see in~§2, that $b$~is
alternating whereas $b_R$~is not.
\begin{proof}[\protect{Proof of Prop.~\ref{prop:trace-local}}]
By the structure theorem of modules over a principal ring, we may
write~$M = ⊕ R_{m_i}$ for some finite sequence of integers~$m_i$. If the
result holds for modules~$(M', N)$ and~$(M'', N)$ then it also holds
for~$(M' ⊕ M'', N)$. Therefore, using induction on the length of both
modules, we only need to prove the case where~$M = R_m$ and~$N = R_n$
where~$m ≥ n$. In this case, since $b$~commutes with~$k[π]$, we see that
for~$x = ∑ x_i π^i$, $y = ∑ y_i π^i$:
\begin{equation}
b(x,y) \;=\; ∑_{i,j} x_i y_j\: b(1, π^{i+j})
\;=\; ∑_{i+j+r = m-1} x_i y_j\: b(1, π^{m-1-r}).
\end{equation}
Let~$c = ∑ b(1, π^{i}) π^{m-1-i}$. Since $b(1, π^{r}) = 0$ for all~$r ≥
n$, $c$~belongs to the ideal~$π^{m-n} R_m = \Hom_R (R_n, R_m)$, so that
the product~$c\,y$ is well-defined in~$R_m$. The bilinear form~$b_R$ is
finally the form defined by~$b_R(x,y) = c\:xy$.
\end{proof}
Assembling the Propositions~\ref{prop:trace-form}
and~\ref{prop:trace-local}, we see that for any local pencil~$(b_{∞},
b_0)$, writing~$R = k[c]$ where $c$~is the characteristic endomorphism
of~$b$, there exists a unique $R$-bilinear form~$b_R$
on~$V$ such that~$b_{∞} = \Tr_{K/k} ∘ τ_d ∘ b_R$.
Local pencils are thus equivalent to $R$-bilinear forms on~$V$.
Moreover, a morphism of pencils~$s: b → b'$ is a
$k$-linear map such that~$\transpose{s} ∘ b_{∞} ∘ s = b'_{∞}$
and~$s^{-1} ∘ c ∘ s = c'$, or in other words a $R$-linear congruence
between the $R$-bilinear forms~$b_R$ and~$b'_R$. Therefore, solving the
local IP1S problem amounts to computing a congruence matrix
between two~$R$-bilinear forms on a $R$-module of finite length.
\note{C'est à peu près ce que dit (22) du preprint IP2S.}
\begin{prop}\label{prop:diag-bigblock}
Let~$M$ be a $R$-module of finite length and $b$~be a $R$-bilinear form
on~$M$. If the $k$-bilinear form~$τ ∘ b$ is regular, then there exists an
orthogonal decomposition~$M = \operp M_d$, where $M_d$~is a finite free
module over~$R_d = R / π^d$ and $b$~is regular, as a
$R_d$-bilinear form, on each module~$M_d$.
\end{prop}
\begin{proof}
We reason by induction on the length of~$R$. By the structure theorem for
modules over the principal ring~$R$, there exists a decomposition~$M = F
⊕ N$ where $F$~is free over~$R = k[π]/π^m$ and~$π^{m-1} N = 0$. In
particular, $N$~is a $R'$-module, where~$R' = R / π^{m-1}$; we define~$τ'
= τ_{m-1}: R' → k$ and note that~$τ'(a) = τ(π a)$ for all~$a ∈ R'$.
We first show that the restriction of~$b$ to~$F$ is regular. Let~$x ∈ F$
such that~$R x$ is a direct factor: this means that~$π^{m-1} x ≠ 0$.
Since $τ ∘ b$~is regular, there exists~$y ∈ M$ such that~$τ(b(π^{m-1} x,
y)) = 1$. This implies that~$b(x, y) ≡ 1 \pmod{π}$. Let~$y = y' + y''$
where~$y' ∈ F$ and~$y'' ∈ M/F$: since~$π^{m-1} y'' = 0$, we have~$π^{m-1}
b(x, y'') = 0$ and therefore~$b(x, y'') ∈ π R$. Therefore, $a = b(x, y')
≡ b(x, y) ≡ 1 \pmod{π}$. In particular, $a ∈ R^{×}$, so that~$b(x, a^{-1}
y') = 1$, which shows that $b$~is regular on~$F$.
Let now~$y ∈ N$. Since $b$~is regular on~$F$, there exists a unique~$f(y)
∈ F$ such that, for all~$x ∈ F$, $b(x, y) = b(x, f(y))$. This implies
that the map~$N → M, y ↦ y - f(y)$ is orthogonal to~$F$, and therefore
defines a $b$-orthogonal decomposition~$M = F ⊕ N$. Finally,
since~$π^{m-1} y = 0$, the map~$b$ has its values in~$π R$, which means
that there exists a $R'$-bilinear form~$b'$ on~$N$ such that~$b = π b'$;
since $τ ∘ b$~is regular, its orthogonal summand $τ' ∘ b'$~is regular,
and we may apply the induction hypothesis to the $R'$-bilinear form~$b'$
on~$N$.
\end{proof}
We shall generally call $b$ \emph{regular} if $τ ∘ b$~is regular as a
$k$-bilinear form. If $M$~is free then this does not conflict with the
standard definition of $R$-regularity.
Note that, if we do not assume $b$~to be regular, then there does not
necessarily exist a decomposition equivalent to that of
Prop.~\ref{prop:diag-bigblock}. For example, the $R_2$-bilinear form
over~$M = R_1 ⊕ R_2$ defined by~$b(x_1 ⊕ x_2, y_1 ⊕ y_2) = (π x_1) y_2 -
x_1 (π y_2)$ is not diagonalizable.
\begin{prop}\label{prop:bilinear-odd}
Assume that $2 ≠ 0$ in~$k$; let~$R = k[π]/π^m$ and let~$M$ be a free
$R$-module.
\begin{enumerate}
\item Any regular $R$-bilinear form~$b$ on~$M$ is diagonalizable.
\item Assume that $k$~is finite. Then any regular $R$-bilinear form~$b$
on~$M$ is equivalent to one of the diagonal forms~$(1, …, 1)$ or~$(1, …,
d)$, where $d ∈ k^{×}$~is not a square.
\end{enumerate}
\end{prop}
\subsection{Quadratic pencils in characteristic two}
If~$2 ≠ 0$ in~$k$, then quadratic pencils are the same as bilinear
pencils. Therefore, we shall only concern ourselves here with the case
where $k$~has characteristic two. We say that a quadratic form
\emph{commutes} with~$R$ if its polar form commutes in the above sense.
\begin{prop} \label{prop:trace-quad}
Let~$K$ be a finite separable extension of~$k$ and~$V$ be a $K$-vector
space. For any $k$-quadratic form~$q: V ⊗ V → k$ commuting with~$K$,
there exists a unique $K$-quadratic form~$q_K: V ⊗ V → K$ such that~$q =
\Tr_{K/k} \, ∘ \, q_K$.
\end{prop}
\note{À peu près la prop. 5 du papier Asiacrypt.}
\begin{proof}
Let~$b$ be the polar form of~$q$. By~\ref{prop:trace-form},
there exists~$b_K ∈ K$
such that~$b(x,y) = \Tr_{K/k} (b_K xy)$; in particular, since $b$~is
alternating, for all~$x$ we have~$\Tr_{K/k} (b_K x^2) = 0$ and
therefore~$b_K = 0$. Therefore, $q$~is a semi-linear form; since the
trace map is non-degenerate, there exists~$q_K$ such that~$q(x) =
\Tr_{K/k} (q_K x^2)$.
As in the bilinear case, the $n$-dimensional case directly follows by
taking coordinates. Here all diagonal entries correspond to quadratic
forms, and all others to bilinear forms; all of these commute with~$K$.
\end{proof}
We again define~$R = R_n = k[π]/π^n$. We recall that the Frobenius
morphism on~$R$ is defined by~$σ(x π^i) = x^2 π^i$; we also define the
Verschiebung morphism by~$V(x π^i) = x π^{2i}$, and note that both
maps~$σ$ and~$V$ are ring morphisms and that their composition, in any
order, is the squaring map. Moreover, we have the decomposition~$R = V(R)
⊕ π V(R)$.
\begin{lem}\label{lem:gamma-polar}
Let~$γ: R → R$ be the application defined, for~$x = V(y) + π V(z)$,
by~$γ(x) = π V(yz)$. Then, for all~$u, x ∈ R$, we have
\begin{enumerate}
\item $γ(ux) = γ(u) x^2 + u^2 γ(x)$;
\item $γ(u+x) = γ(u) + γ(x) + ux + V(α)$ for some~$α ∈ R$.
\end{enumerate}
\end{lem}
\begin{proof}
(i)~Write~$x = V(y) + π V(z)$ and~$u = V(v) + π V(w)$. Noticing
that~$σ(x) = y^2 + π z^2$, we then have
\begin{equation}\label{eq:gamma-prod}
\begin{split}
γ(ux) &= π V\pa{(vy + π wz)(wy+vz)}\\
&= π V\pa{vw (y^2 + π z^2)} + π V \pa{(v^2+πw^2) yz}\\
&= γ(u) V(σx) \;+\; V(σu) γ(x) \quad = γ(u) x^2 + u^2 γ(x).
\end{split}
\end{equation}
(ii) follows from
$γ(u+x) - γ(u) - γ(x) = π V(vz + wy) = ux - V(vy+πwz)$.
\end{proof}
We recall that $τ(xy)$~is a non-degenerate $k$-bilinear form
on~$R$, and we write~$R^{τ}$ for the $τ$-orthogonal of~$V(R)$: this is
the space of~$x$ such that, for all~$y ∈ R$, $τ(xV(y)) = 0$ (or
equivalently, $τ(x y^2) = 0$); this space is linearly spanned by
the~$π^{n-2i}$ for~$0 ≤ i ≤ n/2$. We call the elements of~$R^{τ}$
\emph{$τ$-alternating}, for reasons made obvious by the following
proposition.
\begin{prop}\label{prop:trace-alt}
Let~$M = ⨁ R_{n_i}$ be a $R_m$-module of finite length.
\begin{enumerate}
\item Let~$b$ be the $R$-bilinear form defined by~$b((x_i), (y_i)) = ∑
b_{i,j} x_i y_j$. Then $τ ∘ b$~is alternating if, and only if, $b_{i,j} +
b_{j,i} = 0$ and if the diagonal coefficients $b_{i,i}$ are
$τ$-alternating.
\item Let~$q$ be a $k$-quadratic form on~$M$ commuting with~$R$, $b$~its
polar form, and $b_R$ be the unique $R$-quadratic form such that~$b = τ ∘
b_R$, and~$b_{i,j}$ the coefficients of~$b_R$; then there exists a
$σ$-linear form~$a: M → R$ such that, for all~$x = (x_i) ∈ M$,
\begin{equation}
q(x) \;=\; τ\pa{a(x) + ∑_{i} b_{i,i} γ(x_i) + ∑_{i < j} b_{i,j} x_i x_j}.
\end{equation}
\end{enumerate}
\end{prop}
\begin{proof}
(i) follows from a simple computation. To prove~(ii), we see by
Lemma~\ref{lem:gamma-polar} that for all~$c ∈ R^{τ}$, $τ(c γ(x))$~is a
quadratic form on~$R$ with polar~$τ(c\, xy)$. It follows by linearity
that~$q' = τ(∑ b_{i,i} γ(x_i) + ∑ b_{i,j} x_i x_j)$ is a $k$-quadratic
form on~$M$, which has the same polar as~$q$, so that
the difference~$q - q'$ is a $σ$-linear form.
\end{proof}
% Let~$\Alt_R (M)$ be the $R$-module of alternating bilinear forms and
% $\Alt\kR (M)$ be the $k$-vector space of forms commuting with~$R$. We
% obtain the following commuting diagram, where all lines and columns are
% exact:
% \begin{equation} \xymatrix{
% \Hom_R^{σ} (M, R) \ar[r]\ar@{=}[d] & \Quad_R (M)\ar[r]\ar[d] &
% \Alt_R (M)\ar[d] \\
% \Hom_k^{σ} (M, k) \ar[r]& \Quad\kR (M) \ar[d]\ar[r] &
% \Alt\kR (M) \ar[d] \\
% & \Hom_R^{σ} (M, R^{τ}) \ar@{=}[r] &\Hom_R^{σ} (M, R^{τ}) \\
% }
% \end{equation}
% Here~$\Hom^{σ}$ is the set of $σ$-linear maps.
We now give a standard form for the bilinear form~$b_R$. We note that
this is a symmetric bilinear form, and that its diagonal coefficients are
alternating; if the length~$m$ of~$R$ is even, this means that they are
squares in~$R$, whereas if $m$~is odd, then the diagonal coefficients
belong to~$π V(R)$.
\begin{prop}\label{prop:eqv-norm}
Let~$R = k[π]/π^r$ and let~$M$ be a free $R$-module. For any bilinear
form~$b$ on~$M$, define the \emph{norm} of~$b$ as the ideal~$\fr n b$ of~$R$
generated by the elements~$b(x, x)$ for~$x ∈ M$.
Then two non-degenerate, $τ$-alternating forms are equivalent if and only
if they have the same norm.
\end{prop}
% Let~$R = R_m$ and let~$M$ be a free $R$-module and $b$~be a
% $τ$-alternating, non-degenerate bilinear form on~$M$. Let~$\fr a$ be the
% ideal of~$R$ generated by the values~$b(x, x)$ for~$x ∈ M$.
% \begin{enumerate}
% \item If $\fr a = R$ then $b$~is isomorphic to the bilinear form with
% identity matrix.
% \item Else, for any generator~$a$ of~$\fr a$, $b$~is isomorphic to the
% orthogonal direct sum
% \[ \mat{a & 1\\1&0} ⟂ \mat{0 & 1\\1&0} ⟂ … ⟂ \mat {0 & 1\\ 1 & 0}. \]
% \end{enumerate}
% Furthermore, the first case may occur only if $R$~has even length.
% \end{prop}
\begin{proof}
We show that the bilinear form~$b$ is equivalent to either the form with
identity matrix, if $\fr n b = R$; or the orthogonal direct sum
\[ \mat{a & 1\\1&0} ⟂ \mat{0 & 1\\1&0} ⟂ … ⟂ \mat {0 & 1\\ 1 & 0}, \]
where $a$~is a generator of~$\fr n b$, if $\fr n b ⊂ π R$.
We write~$[u]$ for the diagonal form with coefficient~$u$ and~$H_{v, w}$ for
the form with matrix~$\mat{v&1\\1&w}$. We also write~$H = H_{0,0}$.
By~\cite[§2]{milnor2}, the $k$-bilinear form~$b_k = b ⊗_R k$ has a
decomposition~$b_k ≃ [d_1] ⟂ … ⟂ [d_r] ⟂ H^s$, where all~$d_i$ are
non-zero since $b$~is non-degenerate. We distinguish two cases.
\subparagraph{Case 1: $\fr n b = R$.}
Then the list of~$d_i$ in the above
decomposition is not empty. Moreover, since the field $k$~is perfect, all
elements~$d_i$ are squares in~$k$, so that up to a coordinate change we
may assume that~$d_i = 1$. Finally, we note that the
matrix~$\mat{1&1&0\\1&0&-1\\1&1&-1}$ is an isomorphism between the
forms~$[1] ⟂ [1] ⟂ [-1]$ and~$[1] ⟂ H$, so that if $r ≥ 1$ we may cancel
all the direct factors~$H$ and in this case $b_k$~is isomorphic to the
bilinear form with identity matrix.
By~\cite[Corollary 3.4]{baeza}, the diagonalization of~$b_k$ lifts to a
diagonalization $b ≃ [a_1] ⟂ … ⟂ [a_r]$ over~$R$. Since $b$~is
$τ$-alternating, all coefficients~$a_i$ are $τ$-alternating. If the
length of~$R$ is odd, then the set of $τ$-alternating elements of~$R$
is~$π V(R)$, which contradicts the regularity of~$b$. Therefore, the
length of~$R$ is even, which means that $a_i ∈ V(R)$ and they are
therefore squares in~$R$. From this we deduce that $b$~is isomorphic to
the identity bilinear form.
\subparagraph{Case 2: $\fr n b ⊂ π R$.}
This means that the form~$b_k$ is alternating, so that~$b_k ≃ H^s$.
By~\cite[Corollary 3.4]{baeza}, this decomposition again lifts to a
decomposition $b ≃ H_{u_1, v_1} ⟂ … ⟂ H_{u_s, v_s}$, where all
coefficients~$u_i, v_i$ are $τ$-alternating.
Let $b = H_{u,v}$ be $τ$-alternating and regular; up to a swap of~$u, v$, we
may write it as~$H_{u, u a^2}$ for some~$a ∈ R$. Since $b$~is
regular, $1 + a u ∈ R^{×}$. Therefore, the coordinate
change $P = \mat{1+a u & a/(1+a u) \\ u & 1/(1+a u)}$ transforms the
bilinear form~$b$ to~$H_{u, 0}$.
Finally, we note that the bilinear form~$b = H_{u, 0} ⟂ H_{u a^2 u, 0}$ is
isomorphic to~$H_{v, 0} ⟂ H_{0,0}$ via the coordinate change
% \begin{equation}
$\mat{1 & 0 & a & 0\\0&1&0&0\\0&0&1&0\\au & a & 0 & 1}$.
% \end{equation}
\end{proof}
The \emph{hyperbolic plane} is the module $R^2$, equipped with the
$k$-quadratic form~$q(x) = x_1 x_2$; this form commutes with~$R$. The
corresponding $τ$-alternating bilinear form~$b_R$ has the
matrix~$\mat{0&1\\1&0}$. A quadratic module is \emph{hyperbolic} if it is
isomorphic to the orthogonal sum of copies of the hyperbolic plane.
\begin{prop}
Any free regular $(k, R)$-quadratic module is isomorphic
to the orthogonal direct sum of a hyperbolic space and of a quadratic form
of dimension at most four.
\end{prop}
\begin{proof}
Let~$M = R^m$ and $q$~be a regular quadratic form on~$M$, commuting
with~$R$. Let~$b$ be the polar of~$q$. By Prop.~\ref{prop:eqv-norm}, we
know that $b_R$~is isomorphic to some form
\begin{equation}
\mat{a & 1\\1 & 0} ⟂ \mat{0 & 1\\1 & 0}^{⊗m-2}, a ∈ R^{τ}, \quad\text{or}\quad
\mat{1} ⟂ \mat{0&1\\1&0}^{⊗m-1}.
\end{equation}
For simplicity, we assume the first case; the proof in the second case is
the same.
\end{proof}
\begin{prop}
Any free regular $(k, R)$-quadratic pencil is isomorphic to the
orthogonal direct sum of a hyperbolic pencil and of a quadratic pencil of
dimension at most six.
\end{prop}
\begin{proof}
\note{Pareil...}
\end{proof}
% \begin{lem}\label{lem:exp-binary}
% Let~$E(x)$ be the formal series
% \begin{equation}
% E(x) = ∑_{i ≥ 0} x^{2^{i}-1} \quad ∈ ℤ[[x]].
% \end{equation}
% Then we have the equality in~$\F_2[[x,y]]$:
% \begin{equation}
% \frac{E(x+y)}{E(x)}{E(y)} \;=\; 1 + ∑_{i ≥ 0} ∑_{j=1}^{2^i-1} x^j
% y^{2^i-j}.
% \end{equation}
% \end{lem}
%
%
% \begin{lem}
% For any integer~$m$, we have a group isomorphism
% \begin{equation*}
% R_{2m}^{×} / (R_{2m}^{×})^2 \quad ≃ \quad k^m.
% \end{equation*}
% \end{lem}
%
% \begin{proof}
% For any~$i ≤ n-1$, let~$f_i: k → R_{2m}^{×}$ be the map defined
% by~$f_i(x) = E(π^{2i+1} x)$, where $E$~is the map defined in
% Lemma~\ref{lem:exp-binary}. We then see that $f_i(x+y)/f_i(x) f_i(y)$~is
% a square in~$R_{2m}^{×}$. Since the $m$~maps $f_0, …, f_{m-1}$ are
% obviously linearly independent, they define the required isomorphism.
% \end{proof}
%
Since the bilinear form~$τ$ is non-degenerate, the linear map~$V: R → R$
has an adjoint~$θ$, defined by $θ(π^{n-1-2i}) = π^{n-1-i}$: for all~$x,
y$, we have~$τ(x\, V(y)) = τ(θ(x)\,y)$. A direct computation proves the
following result.
\note{L'application~$θ$ est la même que celle d'Asiacrypt ; la définition
ici comme adjoint du Verschiebung est un peu plus naturelle. Cependant,
la forme canonique n'est pas la même (du tout) que pour Asiacrypt, et
plus facile à manipuler.
}
Let~$q$ be a $k$-quadratic form on~$M$ commuting to~$R$, and write~$q$ as
in Prop.~\ref{prop:trace-alt}; then there exists elements~$a_i ∈ R$ such
that $a((x_i)) = τ(∑ a_i σ(x_i))$. Let~$u: M' → M$ be a $R$-linear map,
and define the $k$-quadratic form~$q' = q ∘ u$ on~$M'$ and the
corresponding elements~$a'_i ∈ R$ and $R$-bilinear form~$b'_R$. Then a
direct computation shows that, writing~$(u_{i,j})$ for the matrix of~$u$,
\begin{equation}\label{eq:theta}
b'_R = b_R ∘ u, \quad
a'_i = ∑_r a_r σ(u_{r,i}) + θ \pa{
∑_r b_{r,r} γ(u_{r,i}) + ∑_{r<s} b_{r,s} u_{r,i} u_{s,i} }.
\end{equation}
\note{ Il y a une erreur dans le Lemme~2 d'Asiacrypt ; considérer par
exemple la forme bilinéaire~$b(x,y) = τ(xy)$ sur~$R_{2m}$, qui est
non-dégénérée et qui n'a pourtant pas de lagrangien stable par
multiplication par~$π$ (les seuls sous-espaces stables étant les idéaux,
soit les~$π^i R$...). Le mieux qu'on puisse faire est de considérer le
morphisme~$V: R_{2m} → R_{2m}$, qui se factorise en fait en~$V: R_{2m} →
R_m → R_{2m}$, et écrire~$R_{2m} = V(R_m) ⊕ π V(R_m)$, ou encore~$R_{2m}
= R_{m}[π_{2m}]/(π_{2m}^2 - π_m)$. Un lagrangien de~$b$ est alors par
exemple~$V(R_m)$, qui n'est pas stable par multiplication par~$π_{2m}$...
Dans la preuve, l'erreur est dans l'écriture de~$K[π] x ⊕ K[π] y$, alors
que ici $x = y$, donc le signe~$⊕$ est faux.
Mais de toute façon la construction donnée ici est plus simple, donc ce
n'est pas trop grave :-)}
The data of a quadratic pencil on~$M$ with characteristic
endomorphism~$π$ is then equivalent to the data of the two quadratic
forms~$b_{∞}$, given by the bilinear form~$b_R$ and a $σ$-linear
form~$a$, and $b_{0}$, given by~$π b_R$ and a $σ$-linear form~$c$.
% For any~$x ∈ R$ written as~$x = V(y) + π V(z)$, we define~$δ(x) =
% σ^{-1} (yz)$. Then for all~$u = V(v) + π V(w)$, we have
% \begin{equation}
% ux \;=\; V(vy \:+\: π\,wz) \:+\: π V(wy \:+\: vz)
% \end{equation}
% so that
% \begin{equation}\label{eq:delta-deriv}
% \begin{split}
% δ(ux) &= σ^{-1} \pa{wv (y^2 + π z^2) \:+\: (v^2 + π w^2) yz}\\
% &= σ^{-1} (vw · σ(x)) \:+\: σ^{-1} (σu · yz)\\
% &= δ(u) x + u δ(x).
% \end{split}
% \end{equation}
%
% \begin{prop}\label{prop:quad-local}
% Let~$q$ be a quadratic form on~$R$, commuting with~$R$.
% Then there exists elements~$a, c ∈ R$, satisfying~$c^2 = 0$, such that
% \begin{equation*}\label{eq:q-sigma-1}
% q(σ^{-1}x) = τ (a x \:+\: c\, δ(x)).
% \end{equation*}
% \end{prop}
%
% \begin{proof}[Proof of Prop.~\ref{prop:quad-local}]
% Let~$b$ be the polar form of~$q$. By Prop.~\ref{prop:trace-local}, there
% exists~$r ∈ R$ such that~$b(x,y) = τ (rxy)$. Moreover, since $b$~is
% alternating, $r$~is orthogonal to all squares in~$R$: for all~$x ∈ R$ we
% have~$τ(rx^2) = 0$ and therefore, for all~$i$, $τ(π^{2i} r) = 0$.
%
% Define~$a = ∑_i π^{n-i-1} q(π^i)$. We then see that $q(x) = τ(a σ(x)) +
% τ( r ∑_{i < j} x_i x_j)$. Write~$x = V(y) + π V(z)$; then in the last
% sum, only the terms of odd degree in~$π$ are not orthogonal to~$r$, and
% their sum is exactly~$π V(yz)$.
%
% Let~$θ: R → R$ be the adjoint map to~$V$, defined by~$τ(x\, θ(y)) =
% τ(V(x)\, y)$ for all~$x, y$. Then $τ( πr\, V(yz)) = τ(θ(πr)\, yz)$.
% Let~$c = θ(π r)$; since~$θ(π^{n-2i-1}) = π^{n-i-1}$, we have $v_R(c) ≥
% n/2$, so that $c^2 = 0$ in~$R$. Since~$δ(σ(x)) = yz$, we can now write
% $q(x) = τ(a σ(x) + c δ(σx))$, from which the proposition follows by
% applying~$σ^{-1}$.
% \end{proof}
%
% Write~$q = [a, c]$ for the quadratic form defined by
% Proposition~\ref{prop:quad-local}. We deduce from \eqref{eq:delta-deriv}
% that, for all~$u ∈ R$, the form~$q(ux)$ is given by
% \begin{equation}\label{eq:q-transform}
% [a, c] ∘ u = [a\,u \:+\: c\, δ(u),\; c\,u].
% \end{equation}
% Solving the cyclic case of IP1S in characteristic two is therefore
% equivalent to solving the above equation in~$u ∈ R$. This is possible in
% polynomial time since $δ$~is contracting.
% We saw in~§2.1 that any regular bilinear pencil on~$V$ is the orthogonal
% sum of pencil of the form $b_{λ} = t a (λ - c)$, where $t$~is any
% prescribed scalar product on~$V$, $a$~is a self-adjoint map relatively
% to~$t$, and $c$~is the characteristic endomorphism of~$b$.
%
% Using the primary factorization of~$c$, we may assume that its minimal
% polynomial is primary, \emph{i.e.} of the form~$p^d$ where $p$~is
% irreducible. In this case, there exists a finite sequence of
% integers~$n_i$ such that~$V$, as a $k[c]$-module, is isomorphic to~$⨁
% k^{n_i} ⊗_{k} k[c]/p(c)^{i}$. Write~$R_n = k[c]/p(c)^n$ and~$K = R_1 =
% k[c]/p(c)$; then $K$~is a finite extension field of~$k$. Consider the equation~$p(y) = 0$ for~$y ∈ R_n =
% k[c]/p(c)^n$; since $p(c)$~generates the maximal ideal~$\fr m$ of~$R_n$,
% $c$~is an approximate solution of the equation, and since $p$~is
% irreducible, the equation is separated. Therefore, by Hensel's lemma,
% there exists a unique solution~$y ∈ R_n$ of~$p(y) = 0$ such that~$y ≡ c
% \pmod{\fr m}$. This defines a $k$-algebra morphism~$K → R_n$, and we see
% that this is an isomorphism between~$R_n$ and the local
% $K$-algebra~$K[π]/π^n$ sending~$p(c)$ to~$π$.
%
% \note{Je viens de prouver, en d'autres termes, quelque chose qui est déjà
% dans le papier : on peut, au lieu d'utiliser la forme normale de
% Frobenius, utiliser celle où les blocs sur-diagonaux sont la matrice
% identité (ce qui est avant 2.14). Au passage, calculer le commutant
% de~$c$ devient trivial : c'est le commutant de~$c$, et il suffit donc de
% demander que tout le monde soit $k[c]$-linéaire...}
%
% \note{Pas besoin du lemme de Hensel pour prouver le résultat qui
% précède : en effet, $(K[π]/π^n, c)$ et~$(V, c)$ sont deux $k$-espaces
% vectoriels équipés d'un endomorphisme cyclique de même polynôme
% caractéristique, et donc forcément isomorphes. Hensel sert à donner une
% version effective.}
%
% Let~$τ$ be the $k$-linear form on~$R_n$ defined in the following way:
% for~$u = ∑ u_i π^i ∈ R_n$, with~$u_i ∈ K$, let~$τ(u) = \Tr_{K/k}
% (u_{n-1})$. Since $p$~is separable, the trace bilinear form~$u, v ↦
% \Tr_{K/k} (uv)$ is non-degenerate on~$K$. Hence the bilinear form~$t$
% defined by~$t(u,v) = τ(uv)$ is non-degenerate on~$R_n$. Moreover, the
% linear map~$c$, corresponding to multiplication by~$c$, is obviously
% self-adjoint for this bilinear form.
%
% \note{Le paragraphe qui précède est une façon différente et peut-être un
% peu plus courte de construire la matrice $T$ de Gilles ; ce n'est
% d'ailleurs pas la même matrice (on a en fait~$T(u, v) = t(u, h v)$ pour
% un certain~$h ∈ K^{×}$). Pour prouver que $t$~est non-dégénéré : soit~$u
% = u_r π^r + u_{r+1} π^{r+1} + …$ avec~$r ≤ d-1$; puisque $\Tr$~est
% non-dégenéré, il existe~$v ∈ K$ tel que~$\Tr (v u_r) = 1$, et on vérifie
% alors que~$t(u, v π^{d-r-1}) = 1$.}
%
%
% \begin{prop}
% Let~$b$ be a bilinear pencil. Assume that $b$~is local, \emph{i.e.} that
% its characteristic polynomial is~$p(c)^d$ where $p$~is irreducible. Then
% there exists a finite sequence of integers~$(n_i)$ such that $(V, b)$~is
% isomorphic, as a $k$-bilinear pencil, to
% \begin{equation}
% (⨁ E_i ⊗_{K} R_i, \quad ⨁ t_i a_i (λ - c)),
% \end{equation}
% where $E_i$~is the space~$K^{n_i}$ equipped with its canonical scalar
% product, $t_i$~is the trace scalar product on~$R_i$ defined above, and
% $a_i$~is an invertible self-adjoint $k[c]$-linear endomorphism of~$E_i
% ⊗_{K} R_i$.
% \end{prop}
%
\section{Old version of §1}
We write~$K_d$ for the Kronecker module of dimension~$2d+1$: its basis
is~$(e_0, …, e_d, f_1, …, f_d)$, with the bilinear pencil~$b_{λ}$ defined
by~$-b_{∞} (e_j, f_{j+1}) = b_{0} (e_j, f_j) = 1$ and~$0$ else. We shall
write~$e_{d,i}$ and~$f_{d,i}$ when needed to prevent ambiguity. We also
write~$K_d^+$ for the space generated by the vectors~$e_i$ and~$K_d^-$
for that generated by the~$f_i$.
Let~$k^n$ be the $n$-dimensional vector space, equipped with its
canonical scalar product and its canonical basis~$(ε_1, …, ε_n)$. Given a
sequence of integers~$(n_d)$, we write~$E_d = k^{n_d}$. The
bilinear module~$K_d^{n_d}$ is isomorphic to~$E_d ⊗ K_d$, with the basis~$ε_i ⊗
e_j, ε_i ⊗ f_j$. We order this basis in the following way: $ε_1 ⊗ e_0, …,
ε_{n_d} ⊗ e_0, …, ε_{n_d} ⊗ e_d, ε_1 ⊗ f_1, …, ε_{n_d} ⊗ f_1, …, ε_{n_d} ⊗ f_d$.
\begin{prop}
Let~$V = ⨁ K_d^{n_d} = ⨁ E_d ⊗ K_d$ be a module with a totally irregular
bilinear pencil. Each automorphism of~$V$ is the product, in
this order, of the following maps:
\begin{enumerate}
\item an endomorphism of the form
\[\begin{array}{rcl}
u ⊗ e_{d,i} & \longmapsto & \displaystyle ∑_{m ≥ 0} ∑_{j=0}^{m}
α_{d,m,j} (u) e_{d-m,i-j},\\
u ⊗ f_{d,i} & \longmapsto & \displaystyle ∑_{m ≥ 0} ∑_{j=0}^{m}
β_{d,m,j} (u) f_{d+m,i+j},
\end{array}\]
where~$α_{m,j}: E_{d} → E_{d-m}$ and~$β_{m,j}: E_{d} →
E_{d+m}$ are $k$-linear maps such that, when writing~$α_{d,m} = ∑ x^j
α_{d,m,j}$ and~$β_{d,m} = ∑ y^j β_{d,m,j}$, we have the relation
\begin{equation}
∑_{i+j=m} \transpose{α_{d+m,i}} β_{d,j} = \begin{cases}
1 & \text{if $m = 0$,}\\0 & \text{if $m ≥ 1$.}
\end{cases}
\end{equation}
\item the map defined by