File: tutorialCongruence.xml

package info (click to toggle)
gap-hap 1.70%2Bds-1
  • links: PTS
  • area: main
  • in suites: forky, sid
  • size: 56,612 kB
  • sloc: xml: 16,139; sh: 216; javascript: 155; makefile: 126; ansic: 47; perl: 36
file content (848 lines) | stat: -rw-r--r-- 48,311 bytes parent folder | download
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
<Chapter><Heading>Congruence Subgroups, Cuspidal Cohomology  and Hecke Operators</Heading>


In this chapter we explain how HAP can be used to make computions 
about modular forms associated to congruence subgroups <M>\Gamma</M> of <M>SL_2(\mathbb Z)</M>. Also, in  Subsection 10.8 onwards, we demonstrate
 cohomology computations for the <E>Picard group</E> <M>SL_2(\mathbb Z[i])</M>,
 some <E>Bianchi groups</E>
 <M>PSL_2({\cal O}_{-d}) </M> where <M>{\cal O}_{d}</M> is the ring of integers of <M>\mathbb Q(\sqrt{-d})</M> for  square free positive integer <M>d</M>, and some other groups of the form
<M>SL_m({\cal O})</M>, 
<M>GL_m({\cal O})</M>,
<M>PSL_m({\cal O})</M>, 
<M>PGL_m({\cal O})</M>,
 for <M>m=2,3,4</M> and certain <M>{\cal O}=\mathbb Z, {\cal O}_{-d}</M>.
 

<Section Label="sec:EichlerShimura"><Heading>Eichler-Shimura isomorphism</Heading>

<P/>We begin by recalling the Eichler-Shimura isomorphism
<Cite Key="eichler"/><Cite Key="shimura"/>
<Display> S_k(\Gamma) \oplus \overline{S_k(\Gamma)} \oplus E_k(\Gamma) \cong_{\sf Hecke} H^1(\Gamma,P_{\mathbb C}(k-2))</Display>
<P/> which relates the cohomology of groups to the theory of modular forms associated to a finite index subgroup <M>\Gamma</M> of <M>SL_2(\mathbb Z)</M>. In subsequent sections we explain how to compute with the right-hand side of the isomorphism. But first,
for completeness,  let us define the terms on the  left-hand side.

<P/> Let <M>N</M> be a positive integer. A subgroup <M>\Gamma</M> of <M>SL_2(\mathbb Z)</M> is said to be a  <E>congruence subgroup</E> of level <M>N </M> if it contains the kernel of the canonical homomorphism  <M>\pi_N\colon SL_2(\mathbb Z) \rightarrow SL_2(\mathbb Z/N\mathbb Z)</M>. 

So any congruence subgroup is of finite index in <M>SL_2(\mathbb Z)</M>, 
but the converse is not true. 

<P/>One congruence subgroup of particular interest is the group
<M>\Gamma_1(N)=\ker(\pi_N)</M>, known as the <E>principal congruence subgroup</E> of level <M>N</M>. Another congruence subgroup of particular interest is the group  <M>\Gamma_0(N)</M> of those matrices that project to upper triangular matrices
in <M>SL_2(\mathbb Z/N\mathbb Z)</M>.  

<P/>A <E>modular form</E> of weight <M>k</M> for a congruence subgroup
<M>\Gamma</M> is a complex valued  function on the upper-half plane,
 <M>f\colon {\frak{h}}=\{z\in \mathbb C : Re(z)>0\} \rightarrow \mathbb C</M>, satisfying:
<List>
<Item> <M>\displaystyle f(\frac{az+b}{cz+d}) = (cz+d)^k f(z)</M> for <M>\left(\begin{array}{ll}a&amp;b\\ c &amp;d \end{array}\right) \in \Gamma</M>,
</Item>
<Item> <M>f</M> is `holomorphic' on the <E>extended upper-half plane</E> 
<M>\frak{h}^\ast = \frak{h} \cup \mathbb Q \cup \{\infty\}</M> obtained from the upper-half plane by `adjoining a point at each cusp'.
</Item>
</List>
The collection of all weight <M>k</M> modular forms for <M>\Gamma</M> form a vector space <M>M_k(\Gamma)</M> over <M>\mathbb C</M>.

<P/>A modular form <M>f</M> is said to be a <E>cusp form</E> if <M>f(\infty)=0</M>. The collection of all weight <M>k</M> cusp forms for <M>\Gamma</M> form a vector subspace <M>S_k(\Gamma)</M>. There is a decomposition 
<Display>M_k(\Gamma) \cong S_k(\Gamma) \oplus E_k(\Gamma)</Display>
<P/> involving a summand <M>E_k(\Gamma)</M>  known as the <E>Eisenstein
space</E>. See <Cite Key="stein"/> for further introductory details on modular forms. 
<P/>The Eichler-Shimura isomorphism is more than an isomorphism of vector spaces. It is an isomorphism of Hecke modules: both sides admit  notions of <E>Hecke operators</E>, and the isomorphism preserves these operators. The bar on the left-hand side of the isomorphism denotes complex conjugation, or <E>anti-holomorphic</E> forms. See
<Cite Key="wieser"/> for a full account of the isomorphism.

<P/><P/>
On the right-hand side of the isomorphism, the <M>\mathbb Z\Gamma</M>-module 
<M>P_{\mathbb C}(k-2)\subset \mathbb C[x,y]</M> denotes the space of homogeneous degree 
<M>k-2</M> polynomials with action of <M>\Gamma</M> given by
<Display>\left(\begin{array}{ll}a&amp;b\\ c &amp;d \end{array}\right)\cdot p(x,y) = p(dx-by,-cx+ay)\ .</Display>
In particular <M>P_{\mathbb C}(0)=\mathbb C</M> is the trivial module.
Below we shall compute with the integral analogue <M>P_{\mathbb Z}(k-2) \subset \mathbb Z[x,y]</M>.


<P/><P/>
In the following sections we  explain how to use the right-hand side of the Eichler-Shimura
isomorphism to compute eigenvalues of the Hecke operators restricted to the 
 subspace        
 <M>S_k(\Gamma)</M> of cusp forms.

</Section>

<Section><Heading>Generators for <M>SL_2(\mathbb Z)</M> and the cubic tree</Heading>

<P/> The matrices 
<M>S=\left(\begin{array}{rr}0&amp;-1\\ 1 &amp;0 \end{array}\right)</M>
and <M>T=\left(\begin{array}{rr}1&amp;1\\ 0 &amp;1 \end{array}\right)</M>
generate <M>SL_2(\mathbb Z)</M> and it is not difficult to devise an algorithm for expressing an arbitrary integer matrix <M>A</M>
of determinant <M>1</M> as a word
in <M>S</M>, <M>T</M> and their inverses. The following illustrates such an algorithm.

<Example>
<#Include SYSTEM "tutex/11.1.txt">
</Example>

It is convenient to introduce the matrix <M>U=ST = \left(\begin{array}{rr}0&amp;-1\\ 1 &amp;1 \end{array}\right)</M>. The matrices <M>S</M> and <M>U</M>
also generate <M>SL_2(\mathbb Z)</M>. In fact we have a free presentation
 <M>SL_2(\mathbb Z)= \langle S,U\, |\, S^4=U^6=1, S^2=U^3 \rangle </M>. 

<P/><P/>
The <E>cubic tree</E> <M>\cal T</M> is a tree (<E>i.e.</E> a <M>1</M>-dimensional  contractible regular  
CW-complex) with countably infinitely many edges in which each vertex has degree <M>3</M>.


We can realize  the cubic tree <M>\cal T</M> by taking the
left cosets of <M>{\cal U}=\langle U\rangle</M> in <M>SL_2(\mathbb Z)</M> as vertices, and joining cosets
<M>x\,{\cal U} </M> and <M>y\,{\cal U}</M> by an edge if, and only if, 
<M>x^{-1}y \in {\cal U}\, S\,{\cal U}</M>. Thus the
vertex <M>\cal U </M> is joined to <M>S\,{\cal U} </M>, 
<M>US\,{\cal U}</M> and <M>U^2S\,{\cal U}</M>. The vertices of this tree
are in one-to-one
correspondence with all reduced words in <M>S</M>, <M>U</M>  and <M>U^2</M> 
that, apart from the identity, end in <M>S</M>.

<P/> From our realization of the cubic tree <M>\cal T</M>
we see that <M>SL_2(\mathbb Z)</M> acts on <M>\cal T</M> in such a way that
each vertex is stabilized by a cyclic subgroup conjugate to <M>{\cal U}=\langle U\rangle</M> and each edge is stabilized by a cyclic subgroup conjugate to
<M>{\cal S} =\langle S \rangle</M>.

<P/> In order to store this action of <M>SL_2(\mathbb Z)</M> on
the cubic tree <M>\cal T</M> we just need to record the following finite amount of information.
<P/>
<Alt Only="HTML">&lt;img src="images/fdsl2.png" align="center" width="350" alt="Information for the cubic tree"/>
</Alt>

</Section>

<Section><Heading>One-dimensional fundamental domains and  
generators for congruence subgroups</Heading>

The modular group <M>{\cal M}=PSL_2(\mathbb Z)</M> is isomorphic, as an abstract group, 
to the free product <M>\mathbb Z_2\ast \mathbb Z_3</M>. By the Kurosh subgroup 
theorem, any finite index subgroup <M>M \subset {\cal M}</M> is isomorphic to the free product 
of finitely many copies of <M>\mathbb Z_2</M>s, <M>\mathbb Z_3</M>s and <M>\mathbb Z</M>s. A
subset <M>\underline x \subset M</M> is  an <E>independent</E> 
 set of subgroup generators if <M>M</M> is the free product of the cyclic subgroups <M>&lt;x &gt;</M> as <M>x</M> runs over <M>\underline x</M>. Let us say 
that a set of elements in <M>SL_2(\mathbb Z)</M> is <E>projectively 
independent</E> if it maps injectively onto an independent set of subgroup generators <M>\underline x\subset {\cal M}</M>.  
The generating set <M>\{S,U\}</M> for <M>SL_2(\mathbb Z)</M> given in the preceding section is projectively independent. 

<P/> We are interested in constructing a set of generators for a given congruence subgroup <M>\Gamma</M>. If a small generating set for <M>\Gamma</M> is required then we should
 aim
to  construct one which is close to being  projectively independent. 

<P/>
 It is useful to    
invoke the following general result which follows from a perturbation result about free <M>\mathbb ZG</M>-resolutons in <Cite Key="ellisharrisskoldberg" Where="Theorem 2"/> and an old observation of John Milnor that a free <M>\mathbb ZG</M>-resolution can be realized as the cellular chain complex of a CW-complex if it can be so realized in low dimensions.

<P/><B>Theorem.</B> Let <M>X</M> be a contractible CW-complex on which a group
<M>G</M> acts by permuting cells. The cellular chain complex <M>C_\ast X</M> 
is a <M>\mathbb ZG</M>-resolution of <M>\mathbb Z</M> which typically is not free. 
Let <M>[e^n]</M> denote the orbit of the n-cell <M>e^n</M> under the action. Let <M>G^{e^n} \le G</M> denote the stabilizer subgroup of <M>e^n</M>, in which group elements are not required to stabilize <M>e^n</M> point-wise. Let <M>Y_{e^n}</M>
denote a contractible CW-complex on which <M>G^{e^n}</M> acts cellularly and freely. Then there exists a contractible CW-complex <M>W</M>
on which <M>G</M> acts cellularly and freely, and in which the orbits of <M>n</M>-cells are labelled by <M>[e^p]\otimes [f^q]</M> where <M>p+q=n</M> and 
<M>[e^p]</M> ranges over the <M>G</M>-orbits of <M>p</M>-cells  in <M>X</M>, <M>[f^q]</M> ranges over the <M>G^{e^p}</M>-orbits of <M>q</M>-cells in <M>Y_{e^p}</M>.
<P/>

<P/>Let <M>W</M> be as in the theorem. Then the quotient CW-complex <M>B_G=W/G</M> is a classifying space for <M>G</M>. Let <M>T</M> denote a maximal tree in the <M>1</M>-skeleton <M>B^1_G</M>. Basic geometric group theory tells us that  the <M>1</M>-cells in <M>B^1_G\setminus T</M> correspond to
a generating set for <M>G</M>.

<P/>
Suppose  we wish to compute a set of generators for a principal congruence subgroup 
<M>\Gamma=\Gamma_1(N)</M>. In the above theorem take <M>X={\cal T}</M> to be the cubic tree, and note that <M>\Gamma</M> acts freely on <M>\cal T</M> and thus that <M>W={\cal T}</M>. 
To determine the  <M>1</M>-cells of <M>B_{\Gamma}\setminus T</M>
 we need to determine a cellular subspace 
<M>D_\Gamma \subset \cal T</M>   whose images under the action of <M>\Gamma</M>  cover <M>\cal T</M> and are  pairwise either disjoint or identical. The subspace <M>D_\Gamma</M> will not be a CW-complex as it won't be closed, but it can be chosen to be connected, 
and hence contractible. We call <M>D_\Gamma</M> a <E>fundamental region</E> for <M>\Gamma</M>. We denote by <M>\mathring D_\Gamma</M> the 
largest CW-subcomplex of <M>D_\Gamma</M>. The vertices of <M>\mathring D_\Gamma</M> are the same as the vertices of <M>D_\Gamma</M>.
Thus <M>\mathring D_\Gamma</M> is a subtree of the cubic tree with <M>|\Gamma|/6</M> vertices. 
 For each vertex <M>v</M> in  the tree
<M>\mathring D_\Gamma</M> define <M>\eta(v)=3 -{\rm degree}(v)</M>. Then the number of  generators for <M> \Gamma </M>  will
 be <M>(1/2)\sum_{v\in \mathring D_\Gamma} \eta(v)</M>. 

 <P/> 
The following commands determine projectively independent 
generators for <M>\Gamma_1(6)</M> and display
<M>\mathring D_{\Gamma_1(6)}</M>. The subgroup <M>\Gamma_1(6)</M> is free on <M>13</M> generators. 

<Example>
<#Include SYSTEM "tutex/11.2.txt">
</Example>
<P/>
<Alt Only="HTML">&lt;img src="images/pctree6.gif" align="center" width="350" alt="Fundamental region in the   cubic tree"/>
</Alt>

<P/>An alternative but very related approach to computing generators  of congruence subgroups of <M>SL_2(\mathbb Z)</M>
 is described in <Cite Key="kulkarni"/>.
<P/>The congruence subgroup <M>\Gamma_0(N)</M> does not act freely on the vertices of <M>\cal T</M>, and so one needs to incorporate
 a generator for the cyclic stabilizer group according to the above theorem. Alternatively, we can replace the cubic tree by a six-fold cover 
<M>{\cal T}'</M> on whose vertex set  <M>\Gamma_0(N)</M> acts freely. 
This alternative approach will produce a  redundant set of generators.
The following commands  display
<M>\mathring D_{\Gamma_0(39)}</M> for a fundamental region in <M>{\cal T}'</M>. They also use the corresponding generating set for <M>\Gamma_0(39)</M>, 
involving <M>18</M> generators, to compute the abelianization <M>\Gamma_0(39)^{ab}= \mathbb Z_2 \oplus \mathbb Z_3^2 \oplus \mathbb Z^9</M>. The abelianization
  shows that any generating set has at least <M>11</M> generators.

<Example>
<#Include SYSTEM "tutex/11.3.txt">
</Example>
<P/>
<Alt Only="HTML">&lt;img src="images/g0tree39.gif" align="center" width="350" alt="Fundamental region in the   cubic tree"/>
</Alt>

<P/> Note that to compute <M>D_\Gamma</M> one only needs to be able to test whether a given matrix lies in <M>\Gamma</M> or not. 
Given an inclusion <M>\Gamma'\subset \Gamma</M> of congruence subgroups, it is straightforward to use the trees <M>\mathring D_{\Gamma'}</M> and <M>\mathring D_{\Gamma}</M> to compute a system of coset representative for <M>\Gamma'\setminus \Gamma</M>.
</Section>

<Section><Heading>Cohomology of congruence subgroups</Heading>

To compute the cohomology <M>H^n(\Gamma,A)</M> of a congruence subgroup 
<M>\Gamma</M> with coefficients in a <M>\mathbb Z\Gamma</M>-module <M>A</M>
 we need to construct <M>n+1</M> terms of a free
<M>\mathbb Z\Gamma</M>-resolution of <M>\mathbb Z</M>. We can do this by first 
using perturbation techniques (as described in <Cite Key="buiellis"/>) to combine the cubic tree with resolutions for
 the cyclic groups of order <M>4</M> and <M>6</M> in order to produce a free
<M>\mathbb ZG</M>-resolution <M>R_\ast</M> for <M>G=SL_2(\mathbb Z)</M>. This resolution is also a free <M>\mathbb Z\Gamma</M>-resolution with each term of rank
<Display>{\rm rank}_{\mathbb Z\Gamma} R_k = |G:\Gamma|\times {\rm rank}_{\mathbb ZG} R_k\ .</Display>
<P/>For congruence subgroups of lowish index in <M>G</M> this resolution suffices to make computations.

<P/>The following commands compute
<Display>H^1(\Gamma_0(39),\mathbb Z) = \mathbb Z^9\ .</Display>


<Example>
<#Include SYSTEM "tutex/11.4.txt">
</Example>

<P/>This  computation establishes that the space <M>M_2(\Gamma_0(39))</M>
of weight <M>2</M> modular forms is of dimension <M>9</M>.


<P/>The following commands show that 
<M>{\rm rank}_{\mathbb Z\Gamma_0(39)} R_1 = 112</M>
but that it is possible to apply `Tietze like' simplifications to <M>R_\ast</M> to obtain a free <M>\mathbb Z\Gamma_0(39)</M>-resolution <M>T_\ast</M>
with <M>{\rm rank}_{\mathbb Z\Gamma_0(39)} T_1 = 11</M>. It is  more efficient to work with <M>T_\ast</M> when making cohomology computations with coefficients in a module <M>A</M> of large rank. 
<Example>
<#Include SYSTEM "tutex/11.5.txt">
</Example>


<P/>The following commands compute
<Display>H^1(\Gamma_0(39),P_{\mathbb Z}(8)) = \mathbb Z_3 \oplus \mathbb Z_6
\oplus \mathbb Z_{168} \oplus \mathbb Z^{84}\ ,</Display>
<Display>H^1(\Gamma_0(39),P_{\mathbb Z}(9)) = \mathbb Z_2 \oplus \mathbb Z_2 .</Display>


<Example>
<#Include SYSTEM "tutex/11.4a.txt">
</Example>

<P/>This computation establishes that the space <M>M_{10}(\Gamma_0(39))</M>
of weight <M>10</M> modular forms is of dimension <M>84</M>, 
and <M>M_{11}(\Gamma_0(39))</M> is of dimension <M>0</M>. 
(There are never any modular forms of odd weight, and so 
<M>M_k(\Gamma)=0</M> for all odd <M>k</M> and any congruence subgroup 
<M>\Gamma</M>.)

<Subsection><Heading>Cohomology with rational coefficients</Heading>
To calculate cohomology <M>H^n(\Gamma,A)</M> with coefficients in a <M>\mathbb Q\Gamma</M>-module <M>A</M>  it suffices to construct a resolution of <M>\mathbb Z</M> by non-free <M>\mathbb Z\Gamma</M>-modules where <M>\Gamma</M> acts  with finite stabilizer groups on each module in the resolution. Computing over <M>\mathbb Q</M> is computationally less expensive than computing over <M>\mathbb Z</M>. The following commands first compute <M>H^1(\Gamma_0(39),\mathbb Q) = 
H_1(\Gamma_0(39),\mathbb Q)= \mathbb Q^9</M>. As a larger example, they then compute 
<M>H^1(\Gamma_0(2^{13}-1),\mathbb Q) =\mathbb Q^{1365}</M> where <M>\Gamma_0(2^{13}-1)</M> has index <M>8192</M> in <M>SL_2(\mathbb Z)</M>.

<Example>
<#Include SYSTEM "tutex/11.4b.txt">
</Example>
</Subsection>
</Section>

<Section><Heading>Cuspidal cohomology</Heading>
To define and compute cuspidal cohomology we  consider the action of 
<M>SL_2(\mathbb Z)</M> on the upper-half plane <M>{\frak h}</M>
given by
<Display>\left(\begin{array}{ll}a&amp;b\\ c &amp;d \end{array}\right) z =
\frac{az +b}{cz+d}\ .</Display>
A standard 'fundamental domain' for this action is the region

<Display>\begin{array}{ll}
 D=&amp;\{z\in {\frak h}\ :\ |z| &gt; 1, |{\rm Re}(z)| &lt; \frac{1}{2}\} \\ 
&amp;
\cup\ \{z\in {\frak h} \ :\ |z| \ge 1, {\rm Re}(z)=-\frac{1}{2}\}\\ 
&amp; \cup\ \{z \in {\frak h}\ :\ |z|=1, -\frac{1}{2} \le {\rm Re}(z) \le 0\}
\end{array}
</Display>
illustrated below.

<P/><Alt Only="HTML">&lt;img src="images/filename-1.png" align="center" width="450" alt="Fundamental domain in the upper-half plane"/>
</Alt>

<P/> The action factors through an action of
<M>PSL_2(\mathbb Z) =SL_2(\mathbb Z)/\langle
\left(\begin{array}{rr}-1&amp;0\\ 0 &amp;-1 \end{array}\right)\rangle</M>.  The images of <M>D</M> under the action of <M>PSL_2(\mathbb Z)</M> 
cover the upper-half plane, and any two images have at most a single point in common. The possible common points are the bottom left-hand corner point which is stabilized by <M>\langle U\rangle</M>, and  the bottom middle point which is stabilized by <M>\langle S\rangle</M>.  

<P/> A congruence subgroup <M>\Gamma</M>  
has a `fundamental domain' <M>D_\Gamma</M> equal to a union of finitely many 
copies of <M>D</M>, one copy for each coset in <M>\Gamma\setminus SL_2(\mathbb Z)</M>. 
The
quotient space <M>X=\Gamma\setminus {\frak h}</M> is not compact, and can be 
compactified in several ways. We are interested in the Borel-Serre 
compactification.
This is a space <M>X^{BS}</M> for which there is an inclusion
<M>X\hookrightarrow X^{BS}</M> and this inclusion is a homotopy equivalence. 
One defines the  <E>boundary</E> <M>\partial X^{BS} = X^{BS} - X</M> and uses the inclusion <M>\partial X^{BS} \hookrightarrow X^{BS} \simeq X</M> to define the cuspidal cohomology group, over the ground ring <M>\mathbb C</M>, as
<Display> H_{cusp}^n(\Gamma,P_{\mathbb C}(k-2)) = \ker (\ H^n(X,P_{\mathbb C}(k-2)) \rightarrow
H^n(\partial X^{BS},P_{\mathbb C}(k-2)) \ ).</Display>
Strictly speaking, this is the definition of  <E>interior cohomology</E> 
<M>H_!^n(\Gamma,P_{\mathbb C}(k-2))</M> which in general contains the
 cuspidal cohomology as a subgroup. However, for congruence subgroups of
<M>SL_2(\mathbb Z)</M> there is equality 
<M>H_!^n(\Gamma,P_{\mathbb C}(k-2)) = H_{cusp}^n(\Gamma,P_{\mathbb C}(k-2))</M>.
<P/> 
Working over <M>\mathbb C</M> has the advantage of 
avoiding the technical issue that <M>\Gamma </M> does not necessarily act freely on <M>{\frak h}</M> since there are points with finite
cyclic stabilizer groups in <M>SL_2(\mathbb Z)</M>. But it has the disadvantage of losing  information about torsion in cohomology. So HAP confronts the issue
by working with a contractible CW-complex 
<M>\tilde X^{BS}</M> on which <M>\Gamma</M> acts freely, and <M>\Gamma</M>-equivariant inclusion
<M>\partial \tilde X^{BS} \hookrightarrow \tilde X^{BS}</M>. The definition of cuspidal cohomology that we use, which coincides with the above definition when working over <M>\mathbb C</M>, is
<Display> H_{cusp}^n(\Gamma,A) = \ker (\ H^n({\rm Hom}_{\, \mathbb  Z\Gamma}(C_\ast(\tilde X^{BS}), A)\,   ) \rightarrow
H^n(\ {\rm Hom}_{\, \mathbb  Z\Gamma}(C_\ast(\tilde \partial X^{BS}), A)\, \ ).</Display>
<P/>The following data is recorded and, using perturbation theory, is  combined
 with free resolutions for <M>C_4</M> and <M>C_6</M> to constuct <M>\tilde X^{BS}</M>.
 
<P/><Alt Only="HTML">&lt;img src="images/filename-2.png" align="center" width="450" alt="Borel-Serre compactified fundamental domain in the upper-half plane"/>
</Alt>

<P/>
The following commands calculate
<Display>H^1_{cusp}(\Gamma_0(39),\mathbb Z) = \mathbb Z^6\ .</Display>
<Example>
<#Include SYSTEM "tutex/11.6.txt">
</Example>
From the Eichler-Shimura isomorphism and the already calculated dimension
of <M>M_2(\Gamma_0(39))\cong \mathbb C^9</M>, we deduce from this cuspidal cohomology that the space 
<M>S_2(\Gamma_0(39))</M> of cuspidal weight <M>2</M> forms is of dimension <M>3</M>,

and the Eisenstein space <M>E_2(\Gamma_0(39))\cong \mathbb C^3</M> is of  dimension <M>3</M>.

<P/>The following commands show that the space <M>S_4(\Gamma_0(39))</M> of cuspidal weight <M>4</M> forms  is of dimension <M>12</M>.

<Example>
<#Include SYSTEM "tutex/11.6a.txt">
</Example>

</Section>

<Section><Heading>Hecke operators on forms of weight 2</Heading>

A congruence subgroup <M>\Gamma \le SL_2(\mathbb Z)</M> and element <M>g\in SL_2(\mathbb Q)</M>  determine the subgroup
<M>\Gamma' = \Gamma \cap g\Gamma g^{-1} </M> and  homomorphisms
<Display> \Gamma\ \hookleftarrow\ \Gamma'\ \ \stackrel{\gamma \mapsto g^{-1}\gamma g}{\longrightarrow}\ \ g^{-1}\Gamma' g\ \hookrightarrow \Gamma\ . </Display> 
These homomorphisms give rise to homomorphisms of cohomology groups
<Display>H^n(\Gamma,\mathbb Z)\ \ \stackrel{tr}{\leftarrow} \ \ H^n(\Gamma',\mathbb Z) \ \ \stackrel{\alpha}{\leftarrow} \ \ H^n(g^{-1}\Gamma' g,\mathbb Z) \ \  \stackrel{\beta}{\leftarrow} H^n(\Gamma, \mathbb Z) </Display>
with <M>\alpha</M>, <M>\beta</M> functorial maps, and <M>tr</M> the transfer map.
We define the composite <M>T_g=tr \circ \alpha \circ \beta\colon H^n(\Gamma, \mathbb Z) \rightarrow H^n(\Gamma, \mathbb Z)</M> to be the <E> Hecke component </E>  determined by <M>g</M>.

<P/>For <M>\Gamma=\Gamma_0(N)</M>,  prime integer <M>p</M> coprime to <M>N</M>, and cohomology degree <M>n=1</M> we define the <E>Hecke operator</E> <M>T_p =T_g</M> 
where <M>g=\left(\begin{array}{cc}1&amp;0\\0&amp;p\end{array}\right)</M>.

	Further details on this description of Hecke operators can be found in
<Cite Key="stein" Where="Appendix by P. Gunnells"/>.

<P/>The following commands compute <M>T_2</M> and <M>T_5</M>

and
 <M>\Gamma=\Gamma_0(39)</M>. The commands also compute the eigenvalues of these two Hecke operators. The final command confirms that <M>T_2</M> and <M>T_5</M> commute. (It is a fact that <M>T_pT_q=T_qT_p</M> for all  <M>p,q</M>.)   

<Example>
<#Include SYSTEM "tutex/11.7.txt">
</Example>

</Section>


<Section><Heading>Hecke operators on forms of weight <M> \ge 2</M></Heading>
	The above definition of Hecke operator <M>T_p</M> for <M>\Gamma=\Gamma_0(N)</M> extends to a Hecke operator
	<M>T_p\colon H^1(\Gamma,P_{\mathbb Q}(k-2) ) \rightarrow H^1(\Gamma,P_{\mathbb Q}(k-2) )</M> for <M>k\ge 2</M>. 
		We  work over the rationals since that is a setting of much interest.

The following commands compute the matrix of <M>T_2\colon H^1(\Gamma,P_{\mathbb Q}(k-2) ) \rightarrow H^1(\Gamma,P_{\mathbb Q}(k-2) )</M> for <M>\Gamma=SL_2(\mathbb Z)</M> and <M>k=4</M>;
<Example>
<#Include SYSTEM "tutex/11.7b.txt">
</Example>
</Section>

<Section><Heading>Reconstructing modular forms from cohomology computations</Heading>

<P/>Given a modular form <M>f\colon {\frak h} \rightarrow \mathbb C</M> associated to a congruence subgroup <M>\Gamma</M>, and given a compact
edge <M>e</M>  in the tessellation of <M>{\frak h}</M> (<E>i.e.</E> an edge in the cubic tree <M>\cal T</M>) arising from the above fundamental domain 
for 
<M>SL_2(\mathbb Z)</M>,  we can evaluate
<Display>\int_e f(z)\,dz \ .</Display> 
In this way we obtain a cochain <M>f_1\colon C_1({\cal T}) \rightarrow \mathbb C</M> in <M>Hom_{\mathbb Z\Gamma}(C_1({\cal T}), \mathbb C)</M>
	representing a cohomology class <M>c(f) \in H^1(\, Hom_{\mathbb Z\Gamma}(C_\ast({\cal T}), \mathbb C)  \,) = H^1(\Gamma,\mathbb C)</M>. The correspondence <M>f\mapsto c(f)</M> underlies the Eichler-Shimura isomorphism.  

	Hecke operators can be used to recover modular forms from cohomology classes.  
	<P/>
		Let <M>\Gamma=\Gamma_0(N)</M>.  The above defined Hecke operators restrict to operators on cuspidal cohomology. On the left-hand side of the Eichler-Shimura isomorphism 
Hecke operators restrict to  operators <M>T_s\colon S_2(\Gamma) \rightarrow S_2(\Gamma)</M> for <M>s\ge 1</M>.


<P/>Consider the function <M>q=q(z)=e^{2\pi i z}</M> which is holomorphic on <M>\mathbb C</M>.
For any modular form <M>f(z) \in M_k(\Gamma)</M> there are numbers <M>a_s</M> such that
<Display>f(z) = \sum_{s=0}^\infty a_sq^s </Display>
for all <M>z\in {\frak h}</M>. The form <M>f</M> is a cusp form if <M>a_0=0</M>.

<P/>  A non-zero
	cusp form <M>f\in S_2(\Gamma)</M> is a cusp <E>eigenform</E> if it is simultaneously an eigenvector for the Hecke operators <M>T_s</M> for all <M>s =1,2,3,\cdots</M> coprime to the level <M>N</M>. A cusp eigenform is said to be <E>normalized</E> if its coefficient <M>a_1=1</M>. It turns out that if <M>f</M> is  normalized  then the coefficient <M>a_s</M> is an eigenvalue for <M>T_s</M> (see for instance <Cite Key="stein"/> for details).
 It can be shown <Cite Key="atkinlehner"/> that <M>S_2(\Gamma_0(N))</M> admits a "basis constructed from eigenforms".

<P/> This all implies that, in principle,
 we can construct an  approximation to an explicit
 basis for the space <M>S_2(\Gamma_0(N))</M> of cusp
 forms by computing eigenvalues for Hecke operators. 

<P/> Suppose that we would like a basis for <M>S_2(\Gamma_0(11))</M>. The following commands first show that <M>H^1_{cusp}(\Gamma_0(11),\mathbb Z)=\mathbb Z\oplus \mathbb Z</M> from which we deduce that <M>S_2(\Gamma_0(11)) =\mathbb C</M> is <M>1</M>-dimensional and thus admits a basis of eigenforms. Then eigenvalues of 
Hecke operators are calculated to establish that the modular form
<Display>f = q -2q^2 -q^3 +2q^4 +q^5 +2q^6 -2q^7 + -2q^9 -2q^{10} + \cdots </Display>
constitutes a basis for <M>S_2(\Gamma_0(11))</M>. 

<Example>
<#Include SYSTEM "tutex/11.7a.txt">
</Example>

<P/>  For a normalized eigenform <M>f=1 + \sum_{s=2}^\infty a_sq^s</M>
the coefficients <M>a_s</M> with <M>s</M> a composite integer can be expressed in terms of the coefficients <M>a_p</M> for prime <M>p</M>.
If <M>r,s</M> are coprime then <M>T_{rs} =T_rT_s</M>.
If <M>p</M> is a prime that is not a divisor of the level <M>N</M> of <M>\Gamma</M> then
<M>a_{p^m} =a_{p^{m-1}}a_p - p a_{p^{m-2}}.</M>
If the prime <M> p</M> divides <M>N</M> then <M>a_{p^m} = (a_p)^m</M>. It thus suffices to compute the coefficients <M>a_p</M> for prime integers <M>p</M> only.

<P/>
The following commands establish that <M>S_{12}(SL_2(\mathbb Z))</M> has a basis
consisting of one cusp eigenform
<P/><M>q - 24q^2 + 252q^3 - 1472q^4 + 4830q^5 - 6048q^6 - 16744q^7 + 84480q^8 - 113643q^9 </M>
<P/><M>- 115920q^{10} + 534612q^{11} - 370944q^{12} - 577738q^{13} + 401856q^{14} + 1217160q^{15} + 987136q^{16}</M>
<P/><M> - 6905934q^{17} + 2727432q^{18} + 10661420q^{19} + ...</M>

<Example>
<#Include SYSTEM "tutex/11.7c.txt">
</Example>

</Section>

<Section><Heading>The Picard group</Heading>
Let us now consider the <E>Picard group</E> <M>G=SL_2(\mathbb Z[ i])</M> and 
its action on <E>upper-half space</E>
<Display>{\frak h}^3 =\{(z,t) \in \mathbb C\times \mathbb R\ |\ t &gt; 0\} \ . </Display>
To describe the action we introduce the symbol <M>j</M> satisfying <M>j^2=-1</M>, <M>ij=-ji</M> and write <M>z+tj</M> instead of <M>(z,t)</M>. The action is given by 
<Display>\left(\begin{array}{ll}a&amp;b\\ c &amp;d \end{array}\right)\cdot (z+tj) \ = \ \left(a(z+tj)+b\right)\left(c(z+tj)+d\right)^{-1}\ .</Display>

Alternatively,  and more explicitly, the action is given by 
<Display>\left(\begin{array}{ll}a&amp;b\\ c &amp;d \end{array}\right)\cdot (z+tj) \ = \  
\frac{(az+b)\overline{(cz+d) } + a\overline c t^2}{|cz +d|^2 + |c|^2t^2} \ +\ 
\frac{t}{|cz+d|^2+|c|^2t^2}\, j
      \ .</Display>

<P/>A standard 'fundamental domain' <M>D</M> for this action is the following region (with some of the boundary points removed).
<Display>
 \{z+tj\in {\frak h}^3\ |\ 0 \le |{\rm Re}(z)| \le \frac{1}{2}, 0\le {\rm Im}(z) \le \frac{1}{2}, z\overline z +t^2 \ge 1\}
</Display>

<Alt Only="HTML">&lt;img src="images/picarddomain.png" align="center" width="350" alt="Fundamental domain for the Picard group"/>
</Alt>
<P/>The four bottom vertices of <M>D</M> are 
<M>a = -\frac{1}{2} +\frac{1}{2}i +\frac{\sqrt{2}}{2}j</M>,
<M>b = -\frac{1}{2}  +\frac{\sqrt{3}}{2}j</M>,
<M>c = \frac{1}{2}  +\frac{\sqrt{3}}{2}j</M>,
<M>d = \frac{1}{2} +\frac{1}{2}i +\frac{\sqrt{2}}{2}j</M>.

<P/>The upper-half space <M>{\frak h}^3</M> can be retracted onto a <M>2</M>-dimensional subspace <M>{\cal T} \subset {\frak h}^3</M>. The space <M>{\cal T}</M>
is a contractible
<M>2</M>-dimensional regular CW-complex, and the action of the Picard group <M>G</M> restricts to a cellular action of <M>G</M> on <M>{\cal T}</M>. 

<P/>Using perturbation techniques, the <M>2</M>-complex <M>{\cal T}</M> can be combined with free resolutions for the cell stabilizer groups to contruct a regular CW-complex <M>X</M> on which the Picard group <M>G</M> acts freely. 
The following commands compute the first few terms of the free <M>\mathbb ZG</M>-resolution 
  <M>R_\ast =C_\ast X</M>. 
Then   <M>R_\ast</M> is used to compute 
<Display>H^1(G,\mathbb Z) =0\ ,</Display>
<Display>H^2(G,\mathbb Z) =\mathbb Z_2\oplus \mathbb Z_2\ ,</Display>
<Display>H^3(G,\mathbb Z) =\mathbb Z_6\ ,</Display>
<Display>H^4(G,\mathbb Z) =\mathbb Z_4\oplus \mathbb Z_{24}\ ,</Display>
 and compute a free presentation for <M>G</M> involving four generators and seven relators.

<Example>
<#Include SYSTEM "tutex/11.8.txt">
</Example>

We can also compute the cohomology of <M>G=SL_2(\mathbb Z[i])</M> with 
coefficients in a module such as the module 
<M>P_{\mathbb Z[i]}(k)</M> of degree <M>k</M>
homogeneous polynomials with coefficients in <M>\mathbb Z[i]</M>
and with the action described above. For instance, 
the following commands compute
<Display>H^1(G,P_{\mathbb Z[i]}(24)) = (\mathbb Z_2)^4 \oplus \mathbb Z_4
\oplus \mathbb Z_8 \oplus \mathbb Z_{40} \oplus \mathbb Z_{80}\, ,</Display>
<Display>H^2(G,P_{\mathbb Z[i]}(24)) = (\mathbb Z_2)^{24} \oplus \mathbb Z_{520030}\oplus \mathbb Z_{1040060} \oplus  \mathbb Z^2\, ,</Display>
<Display>H^3(G,P_{\mathbb Z[i]}(24)) = (\mathbb Z_2)^{22} \oplus \mathbb Z_{4}\oplus (\mathbb Z_{12})^2 \, .</Display>

<Example>
<#Include SYSTEM "tutex/11.9.txt">
</Example>

</Section>

<Section><Heading>Bianchi groups</Heading>
The <E>Bianchi groups</E> are the groups <M>G=PSL_2({\cal O}_{-d})</M> where <M>d</M> is a square free positive integer and <M>{\cal O}_{-d}</M> is the ring of integers
of the imaginary quadratic field <M>\mathbb Q(\sqrt{-d})</M>. More explicitly, 
<Display>{\cal O}_{-d} = \mathbb Z\left[\sqrt{-d}\right]~~~~~~~~ {\rm if~} d \equiv 1,2 {\rm ~mod~} 4\, ,</Display>
<Display>{\cal O}_{-d} = \mathbb Z\left[\frac{1+\sqrt{-d}}{2}\right]~~~~~ {\rm if~} d \equiv 3 {\rm ~mod~} 4\, .</Display>
These groups act on upper-half space <M>{\frak h}^3</M> in the same way as the Picard group.   Upper-half space can be tessellated by a  'fundamental domain'
 for this action. Moreover, as with the Picard group, this tessellation contains a <M>2</M>-dimensional cellular subspace <M>{\cal T}\subset {\frak h}^3</M>
where <M>{\cal T}</M> is a contractible CW-complex on which <M>G</M> acts cellularly. It should be mentioned that the fundamental domain and the contractible <M>2</M>-complex  <M>{\cal T}</M> are not uniquely determined by <M>G</M>.
 Various algorithms exist for computing <M>{\cal T}</M> and its cell stabilizers. One algorithm due to  Swan 
<Cite Key="swan"/> has been implemented by Alexander Rahm 
<Cite Key="rahmthesis"/> and the output for various values of <M>d</M> are stored in HAP. Another approach is to use Voronoi's theory of perfect forms. This approach has been implemented by Sebastian Schoennenbeck <Cite Key="schoennenbeck"/> and, again, its output for various values of <M>d</M> are 
stored in HAP. The following commands  combine data from Schoennenbeck's algorithm with free resolutions for cell stabiliers to compute
 
<Display>H^1(PSL_2({\cal O}_{-6}),P_{{\cal O}_{-6}}(24)) =
(\mathbb Z_2)^4
\oplus \mathbb Z_{12}
\oplus \mathbb Z_{24}
\oplus \mathbb Z_{9240}
\oplus \mathbb Z_{55440}
\oplus \mathbb Z^4\,,
</Display>
<Display>H^2(PSL_2({\cal O}_{-6}),P_{{\cal O}_{-6}}(24)) =
\begin{array}{l}
(\mathbb Z_2)^{26}
\oplus \mathbb (Z_{6})^8
\oplus \mathbb (Z_{12})^{9}
\oplus \mathbb Z_{24}
\oplus (\mathbb Z_{120})^2
\oplus (\mathbb Z_{840})^3\\
\oplus \mathbb Z_{2520}
\oplus (\mathbb Z_{27720})^2
\oplus (\mathbb Z_{24227280})^2
\oplus (\mathbb Z_{411863760})^2\\
\oplus \mathbb Z_{2454438243748928651877425142836664498129840}\\
\oplus \mathbb Z_{14726629462493571911264550857019986988779040}\\
\oplus \mathbb Z^4\end{array}\ ,
</Display>
<Display>H^3(PSL_2({\cal O}_{-6}),P_{{\cal O}_{-6}}(24)) =
(\mathbb Z_2)^{23}
\oplus \mathbb Z_{4}
\oplus (\mathbb Z_{12})^2\ .
</Display>
Note that the action of <M>SL_2({\cal O}_{-d})</M> on <M>P_{{\cal O}_{-d}}(k)</M> induces an action of <M>PSL_2({\cal O}_{-d})</M>
provided <M>k</M> is even.

<Example>
<#Include SYSTEM "tutex/11.10.txt">
</Example>

<P/>We can also consider the coefficient module
<Display> P_{{\cal O}_{-d}}(k,\ell) = P_{{\cal O}_{-d}}(k) \otimes_{{\cal O}_{-d}} \overline{P_{{\cal O}_{-d}}(\ell)}   </Display>
where the bar denotes a twist in the action obtained from complex conjugation. For an action of the projective linear group we must insist that <M>k+\ell</M> is even. The following commands compute
<Display>H^2(PSL_2({\cal O}_{-11}),P_{{\cal O}_{-11}}(5,5)) = 
(\mathbb Z_2)^8 \oplus \mathbb Z_{60} \oplus (\mathbb Z_{660})^3 \oplus \mathbb Z^6\,,      </Display>
a computation which was first made, along with many other cohomology computationsfor Bianchi groups, by Mehmet Haluk Sengun <Cite Key="sengun"/>.
<Example>
<#Include SYSTEM "tutex/11.10b.txt">
</Example>

<P/>The function <Code>ResolutionPSL2QuadraticIntegers(-d,n)</Code> relies on a limited data base produced by the algorithms implemented by Schoennenbeck and Rahm.  
The function also covers some cases covered by entering a sring
"-d+I" as first variable. These cases  
correspond to projective special groups of module automorphisms of lattices of rank 2 over the integers of the imaginary quadratic number field <M>\mathbb Q(\sqrt{-d})</M> with non-trivial Steinitz-class. In the case of a larger class group there are cases labelled "-d+I2",...,"-d+Ik" and the Ij together with O-d form a system of representatives of elements of the class group modulo squares and Galois action.


 For instance,
the following commands compute 
<Display>H_2(PSL({\cal O}_{-21+I2}),\mathbb Z) = \mathbb Z_2\oplus  \mathbb Z^6\, .</Display>

<Example>
<#Include SYSTEM "tutex/11.11.txt">
</Example>
</Section>

<Section><Heading>(Co)homology of Bianchi groups and <M>SL_2({\cal O}_{-d})</M></Heading>
The (co)homology of Bianchi groups has been studied in papers such as 
		<Cite Key="Schwermer"/>
<Cite Key="Vogtmann"/>	
	<Cite Key="Berkove00"/>
		 <Cite Key="Berkove06"/>
			 <Cite Key="Rahm11"/>
				 <Cite Key="Rahm13"/>
					 <Cite Key="Rahm13a"/>
					 <Cite Key="Rahm20"/>.
						 Calculations in these papers can often be verified by computer. For instance, the calculation
<Display>H_q(PSL_2({\cal O}_{-15}),\mathbb Z) =
	\left\{\begin{array}{ll}
	\mathbb Z^2 \oplus \mathbb Z_6  &amp; q=1,\\
	\mathbb Z \oplus \mathbb Z_6 &amp; q=2,\\
	\mathbb Z_6 &amp; q\ge 3\\
\end{array}\right. 
</Display>
	obtained in <Cite Key="Rahm11"/> can be verified as follows, once we note that Bianchi groups have virtual cohomological dimension 2 and, if all stabilizer groups are periodic with period dividing m, then the homology has period dividing m in degree <M>\ge 3</M>.

		<Example>
<#Include SYSTEM "tutex/11.22.txt">
</Example>

	All finite subgroups of <M>SL_2({\cal O}_{-d})</M> are periodic. Thus the above example can be adapted from PSL to SL for any square=free <M>d\ge 1</M>. For example, the calculation
<Display>H^q(SL_2({\cal O}_{-2}),\mathbb Z) =
        \left\{\begin{array}{ll}
        \mathbb Z  &amp; q=1,\\
        \mathbb Z_6 &amp; q=2 {\rm \ mod\ } 4,\\
	\mathbb Z_2 \oplus \mathbb Z_{12} &amp; q= 3 {\rm \ mod\ } 4\\
	\mathbb Z_2 \oplus \mathbb Z_{24} &amp; q= 0 {\rm \ mod\ } 4  (q>0)\\
	\mathbb Z_{12}  &amp; q= 1 {\rm \ mod\ } 4  (q > 1)\\
\end{array}\right.
</Display>
	obtained in <Cite Key="Schwermer"/> can be verified as follows.



	<Example>
<#Include SYSTEM "tutex/11.23.txt">
</Example>

	 A quotient of a periodic group by a central subgroup of order 2 need not be periodic. For this reason the (co)homology of PSL can be a bit more tricky than SL.
                For example, the calculation
                <Display>H^q(PSL_2({\cal O}_{-13}),\mathbb Z) =
        \left\{\begin{array}{ll}
        \mathbb Z^3 \oplus (\mathbb Z_2)^2  &amp; q=1,\\
        \mathbb Z^2 \oplus \mathbb Z_4 \oplus (\mathbb Z_3)^2 \oplus \mathbb Z_2 &amp; q=2,\\
        (\mathbb Z_2)^q \oplus (\mathbb Z_{3})^2 &amp; q= 3 {\rm \ mod\ } 4\\
        (\mathbb Z_2)^q  &amp; q= 0 {\rm \ mod\ } 4  (q>0)\\
                        (\mathbb Z_2)^q  &amp; q= 1 {\rm \ mod\ } 4  (q>1)\\
        (\mathbb Z_2)^q \oplus (\mathbb Z_{3})^2  &amp; q= 2 {\rm \ mod\ } 4  (q > 2)\\
\end{array}\right.
</Display>
	was obtained in <Cite Key="Rahm11"/>. The following commands verify the calculation in the first 34 degrees, but for a proof valid for all degrees one needs to analyse the computation to spot that there is a certain "periodicity of period 2" in the computations for <M>q\ge 3</M>. This analysis is done in <Cite Key="Rahm11"/>.

		<Example>
<#Include SYSTEM "tutex/11.24.txt">
</Example>

	The Lyndon-Hochschild-Serre spectral sequence  <M>H_p(G/N,H_q(N,A)) \Rightarrow H_{p+q}(G,A)</M> for the groups <M>G=SL_2({\mathcal O}_{-d})</M> and <M>N\cong C_2</M> the central subgroup with <M>G/N\cong PSL_2({\mathcal O}_{-d})</M>, and the trivial module 
	<M>A = \mathbb Z_{\ell}</M>, implies that for primes <M>\ell>2</M> 
		 we have a natural isomorphism <M>H_n(PSL_2({\mathcal O}_{-d}),\mathbb Z_{\ell}) \cong H_n(SL_2({\mathcal O}_{-d}),\mathbb Z_{\ell})</M>. It follows that we have an isomorphism of <M>\ell</M>-primary 
		parts <M>H_n(PSL_2({\mathcal O}_{-d}),\mathbb Z)_{(\ell)} \cong H_n(SL_2({\mathcal O}_{-d}),\mathbb Z)_{(\ell)}</M>.
			Since <M>  H_n(SL_2({\mathcal O}_{-d}),\mathbb Z)_{(\ell)}</M> is periodic in degrees <M>  \ge 3</M> we can recover
				the <M>3</M>-primary part of <M>H_n(PSL_2({\mathcal O}_{-13}),\mathbb Z)</M> in all degrees <M>q\ge1</M>
					from the following computation by ignoring all  <M>2</M>-power factors in the output. 

						 <Example>
<#Include SYSTEM "tutex/11.25.txt">
</Example>

	The ring <M>{\mathcal O}_{-163}</M> is an example of a
		principal ideal domain that is not a Euclidean domain. It seems that no complete calculation of <M>H_n(PSL_2({\mathcal O}_{-163}),\mathbb Z)</M>
			is yet available in the literature. The following comands compute this homology in the first <M>31</M> degrees. The computation suggests a general formula in higher degrees. All but two of the stabilizer groups for
				the action of
			<M>PSL_2({\mathcal O}_{-163})</M> are periodic. The non-periodic group <M>A_4</M> occurs twice in degree <M>0</M>.



				<Example>
<#Include SYSTEM "tutex/11.26.txt">
</Example>

			 	</Section>

<Section><Heading>Some other infinite matrix groups</Heading>

Analogous to the functions for Bianchi groups, HAP has functions
<List>
<Item><Code>ResolutionSL2QuadraticIntegers(-d,n)</Code> </Item>
<Item><Code>ResolutionSL2ZInvertedInteger(m,n)</Code></Item>
<Item><Code>ResolutionGL2QuadraticIntegers(-d,n)</Code></Item>
<Item><Code>ResolutionPGL2QuadraticIntegers(-d,n)</Code></Item>
<Item><Code>ResolutionGL3QuadraticIntegers(-d,n)</Code></Item>
<Item><Code>ResolutionPGL3QuadraticIntegers(-d,n)</Code></Item>
</List>
for computing free resolutions for certain values of <M>SL_2({\cal O}_{-d})</M>,
<M>SL_2(\mathbb Z[\frac{1}{m}])</M>,
<M>GL_2({\cal O}_{-d})</M> and <M>PGL_2({\cal O}_{-d})</M>.
Additionally, the function 
<List>
<Item><Code>ResolutionArithmeticGroup("string",n)</Code></Item>
</List>
 can be used to compute resolutions for groups whose data (provided by Sebastian Schoennenbeck, Alexander Rahm and Mathieu Dutour) is stored in the directory <Code>gap/pkg/Hap/lib/Perturbations/Gcomplexes</Code> .


<P/>For instance, the following commands   compute

<Display>H^1(SL_2({\cal O}_{-6}),P_{{\cal O}_{-6}}(24)) =
(\mathbb Z_2)^4
\oplus \mathbb Z_{12}
\oplus \mathbb Z_{24}
\oplus \mathbb Z_{9240}
\oplus \mathbb Z_{55440}
\oplus \mathbb Z^4\,,
</Display>
<Display>H^2(SL_2({\cal O}_{-6}),P_{{\cal O}_{-6}}(24)) =
\begin{array}{l}
(\mathbb Z_2)^{26}
\oplus \mathbb (Z_{6})^7
\oplus \mathbb (Z_{12})^{10}
\oplus \mathbb Z_{24}
\oplus (\mathbb Z_{120})^2
\oplus (\mathbb Z_{840})^3\\
\oplus \mathbb Z_{2520}
\oplus (\mathbb Z_{27720})^2
\oplus (\mathbb Z_{24227280})^2
\oplus (\mathbb Z_{411863760})^2\\
\oplus \mathbb Z_{2454438243748928651877425142836664498129840}\\
\oplus \mathbb Z_{14726629462493571911264550857019986988779040}\\
\oplus \mathbb Z^4\end{array}\ ,
</Display>
<Display>H^3(SL_2({\cal O}_{-6}),P_{{\cal O}_{-6}}(24)) =
(\mathbb Z_2)^{58}
\oplus (\mathbb Z_{4})^4
\oplus (\mathbb Z_{12})\ .
</Display>


<Example>
<#Include SYSTEM "tutex/11.10a.txt">
</Example>

<P/>The following commands construct free resolutions up to degree 5 for the groups <M>SL_2(\mathbb Z[\frac{1}{2}])</M>,
<M>GL_2({\cal O}_{-2})</M>, <M>GL_2({\cal O}_{2})</M>, <M>PGL_2({\cal O}_{2})</M>, <M>GL_3({\cal O}_{-2})</M>, <M>PGL_3({\cal O}_{-2})</M>.
The final command constructs a free resolution up to degree 3 for <M>PSL_4(\mathbb Z)</M>.
<Example>
<#Include SYSTEM "tutex/11.12.txt">
</Example>

</Section>

<Section><Heading>Ideals and finite quotient groups</Heading>

The following commands first construct the number field <M>\mathbb Q(\sqrt{-7})</M>,
its ring of integers <M>{\cal O}_{-7}={\cal O}(\mathbb Q(\sqrt{-7}))</M>, and the principal ideal <M>I=\langle 5 + 2\sqrt{-7}\rangle \triangleleft {\cal O}(\mathbb Q(\sqrt{-7}))</M>
of norm <M>{\cal N}(I)=53</M>. The  ring <M>I</M> is prime since its norm is a prime number. The primality of <M>I</M> is also demonstrated by observing that the quotient ring  <M>R={\cal O}_{-7}/I</M> is  an integral domain and hence 
isomorphic to the unique finite field of order <M>53 </M>, 
<M>R\cong \mathbb Z/53\mathbb Z</M> . (In a ring of quadratic integers <E>prime ideal</E> is the same as <E>maximal ideal</E>). 

<P/>The finite group <M>G=SL_2({\cal O}_{-7}\,/\,I)</M> is then
 constructed and confirmed to be isomorphic to <M>SL_2(\mathbb Z/53\mathbb Z)</M>. The group <M>G</M> is shown to admit a periodic <M>\mathbb ZG</M>-resolution of <M>\mathbb Z</M> of period dividing <M>52</M>. 
<P/>Finally the integral homology
<Display>H_n(G,\mathbb Z) = \left\{\begin{array}{ll}
0 &amp; n\ne 3,7, {\rm~for~} 0\le n \le 8,\\
\mathbb Z_{2808} &amp; n=3,7,
\end{array}\right.</Display> 
is computed.
<Example>
<#Include SYSTEM "tutex/11.13.txt">
</Example>

<P/>The following commands show that the rational prime <M>7</M>
is not prime in  <M>{\cal O}_{-5}={\cal O}(\mathbb Q(\sqrt{-5}))</M>. Moreover,
 <M>7</M> totally splits in <M>{\cal O}_{-5}</M> since the final command
shows that only    the rational primes <M>2</M> and  <M>5</M> ramify in 
<M>{\cal O}_{-5}</M>.

<Example>
<#Include SYSTEM "tutex/11.14.txt">
</Example>

<P/> For <M>d &lt; 0</M>  the  rings <M>{\cal O}_d={\cal O}(\mathbb Q(\sqrt{d}))</M> are unique factorization domains for precisely
<Display> d = -1, -2, -3, -7, -11, -19, -43, -67, -163.</Display>
This result was conjectured by Gauss, and   essentially proved by Kurt Heegner,
and then later proved by Harold Stark. 

<P/>The following commands construct the classic  example of a prime ideal <M>I</M> that is not principal. They then illustrate reduction modulo <M>I</M>.
<Example>
<#Include SYSTEM "tutex/11.15.txt">
</Example>

</Section>

<Section><Heading>Congruence subgroups for ideals</Heading>

<P/> Given a ring of integers <M>{\cal O}</M> and ideal
<M>I \triangleleft {\cal O}</M> there is a canonical homomorphism
<M>\pi_I\colon SL_2({\cal O}) \rightarrow SL_2({\cal O}/I)</M>. A
subgroup <M>\Gamma \le SL_2({\cal O})</M> is said to be a
<E>congruence subgroup</E> if it contains <M>\ker \pi_I</M>. Thus congruence subgroups are of finite index.  
 Generalizing the definition in <Ref Sect="sec:EichlerShimura"/> above, we define the <E>principal congruence subgroup</E> <M>\Gamma_1(I)=\ker \pi_I</M>, and the congruence subgroup
<M>\Gamma_0(I)</M> consisting of preimages of the upper triangular matrices in
<M>SL_2({\cal O}/I)</M>.

<P/> The following commands construct <M>\Gamma=\Gamma_0(I)</M> for the ideal
<M>I\triangleleft {\cal O}\mathbb Q(\sqrt{-5})</M> generated by <M>12</M>
and <M>36\sqrt{-5}</M>. The group <M>\Gamma</M>
has index <M>385</M> in <M>SL_2({\cal O}\mathbb Q(\sqrt{-5}))</M>.
The final command displays a tree in a Cayley graph for <M>SL_2({\cal O}\mathbb Q(\sqrt{-5}))</M> whose nodes represent a transversal for <M>\Gamma</M>.

<Example>
<#Include SYSTEM "tutex/11.17.txt">
</Example>
<Alt Only="HTML">&lt;img src="images/treesqrt5.gif" align="center" width="350" alt="Information for the cubic tree"/>
</Alt>

<P/>The next commands first construct the congruence subgroup 
<M>\Gamma_0(I)</M> of index <M>144</M> in <M>SL_2({\cal O}\mathbb Q(\sqrt{-2}))</M> for the ideal <M>I</M> in <M>{\cal O}\mathbb Q(\sqrt{-2})</M> generated by
<M>4+5\sqrt{-2}</M>.
 The commands then compute 
<Display>H_1(\Gamma_0(I),\mathbb Z) = \mathbb Z_3 \oplus \mathbb Z_6 \oplus \mathbb Z_{30} \oplus \mathbb Z^8\, ,</Display>
<Display>H_2(\Gamma_0(I), \mathbb Z) = (\mathbb Z_2)^9 \oplus \mathbb Z^7\, ,</Display>
<Display>H_3(\Gamma_0(I), \mathbb Z) = (\mathbb Z_2)^9 \, .</Display>

<Example>
<#Include SYSTEM "tutex/11.16.txt">
</Example>

</Section>

<Section><Heading>First homology</Heading>
The isomorphism <M>H_1(G,\mathbb Z) \cong G_{ab}</M> allows for the computation of first integral homology using computational methods for finitely presented groups. Such methods underly the following computation of
<Display>H_1( \Gamma_0(I),\mathbb Z) \cong \mathbb Z_2 \oplus  \cdots \oplus
\mathbb Z_{4078793513671}</Display>
where <M>I</M> is the prime ideal in the Gaussian integers generated by  <M>41+56\sqrt{-1}</M>.

<Example>
<#Include SYSTEM "tutex/11.18.txt">
</Example>

<P/>We write <M>G^{ab}_{tors}</M> to denote the maximal finite
summand of the first homology group of <M>G</M> and refer to this as the
<E>torsion subgroup</E>.
 Nicholas Bergeron and Akshay Venkatesh   <Cite Key="bergeron"/> 
have conjectured relationships between 
the  torsion in congruence subgroups <M>\Gamma</M> and the volume of their quotient manifold
<M>{\frak h}^3/\Gamma</M>. For instance, for the Gaussian integers
they conjecture
<Display> \frac{\log |\Gamma_0(I)_{tors}^{ab}|}{{\rm Norm}(I)} \rightarrow \frac{\lambda}{18\pi},\ \lambda =L(2,\chi_{\mathbb Q(\sqrt{-1})}) = 1 -\frac{1}{9} + \frac{1}{25} - \frac{1}{49}  + \cdots</Display>
as the norm of the prime ideal <M>I</M> tends to <M>\infty</M>.
The following approximates <M>\lambda/18\pi = 0.0161957</M> and
<M>\frac{\log |\Gamma_0(I)_{tors}^{ab}|}{{\rm Norm}(I)} = 0.0210325</M>
 for the above example.

<Example>
<#Include SYSTEM "tutex/11.19.txt">
</Example>



<P/> The link with volume is given by the Humbert volume formula
<Display>   
{\rm Vol} ( {\frak h}^3 / PSL_2( {\cal O}_{d} ) )
 = \frac{|D|^{3/2}}{24} \zeta_{ \mathbb Q( \sqrt{d} ) }(2)/\zeta_{\mathbb Q}(2)                           </Display>
 valid for square-free <M>d&lt;0</M>, where <M>D</M> is the discriminant of <M>\mathbb Q(\sqrt{d})</M>. The volume of a finite index subgroup <M>\Gamma</M>
	 is obtained by multiplying the right-hand side by the  index <M>|PSL_2({\cal O}_d)\,:\, \Gamma|</M>.

<P/>
	 The following commands produce a graph of 
 <M> \frac{\log |\Gamma_0(I)_{tors}^{ab}|}{{\rm Norm}(I)}</M>  against 
 <M>{\rm Norm}(I)</M> for prime ideals <M>I</M> of norm <M>49 \le {\rm Norm}(I) \le 4357</M> (where one ideal for each norm is taken).

<Example>
<#Include SYSTEM "tutex/11.21.txt">
</Example>

<Alt Only="HTML">&lt;img src="images/primesTorsion.png" align="center" width="1200" alt="graph of torsion versus norm"/>
</Alt>


</Section>
</Chapter>