-
Notifications
You must be signed in to change notification settings - Fork 0
/
Copy pathloekker.tex
1119 lines (1046 loc) · 46.7 KB
/
loekker.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
\chapter{Loops and limits}
\label{chap:loekker}
We now return to the general case. We want to show that the limit
space
$X_{m,\infty}$ is equivalent to the space of loops in the flag
manifold $\GLB{m}$.
\begin{definition}
The loop space $\Omega = \Omega(\GLB{m})$ is the space
\[ \Omega = \set{\g : [0,1] \to \GLB{m} \mid \g(0) = \g(1) =
\Id} \]
with the compact-open topology. This was briefly described in
Section \ref{sec:ls}. Since $\GLB{m}$ is a Riemannian manifold, the
loop space becomes a metric space where loops are close if they are
close at every point.
\end{definition}
\section{The map}
As has been done before, we will start by considering $Y_{m,n}$ and
then quotient out with the torus $\T^n$ afterwards.
For $A = (a_1,\dots,a_n)\in Y_{m,n}$ we
can define a path $ \g_A$ in $\GL_m(\C)$ by starting with
$\g_A(0) = \Id$. Since the matrix $(e_2,\dots,e_m,a_1)$ is
invertible by the definition of $Y_{m,n}$, we know that $a_1$ is
linearly independent of $e_2,\dots,e_m$. Hence it can be written as
\[ a_1 = \sum_{i=1}^m c_i e_i, \]
with $c_1 \neq 0$. Rewriting the expression leads to the formula
\[ \frac{1}{c_1} a_1= e_1 + \frac{1}{c_1} \left(\sum_{i=2}^m c_i
e_i\right). \]
This gives us a path in $\GL_m(\C)$ from the identity to
$\left(\frac{1}{c_1} a_1,e_2,\dots,e_m\right)$:
\[ [0,1]\ni t \mapsto \left( e_1 + t\left(\sum_{i=2}^m
\frac{c_i}{c_1}e_i\right),e_2,\dots,e_m \right) \in \GL_m(\C). \]
This process can be continued, since we can rewrite the similar
expression for $a_2$ as
\[ \frac{1}{d_2}a_2 = e_2 + \frac{1}{d_2}\left(\sum_{i=3}^m d_i
e_i\right) + \frac{d_1}{d_2c_1}a_1, \]
and get a path
\[ t\mapsto \left(\frac{1}{c_1}a_1,e_2+t\left(\sum_{i=3}^m
\frac{d_i}{d_2}e_i +
\frac{d_1}{d_2c_1}a_1\right),e_3,\dots,e_m\right) \]
that starts where the other path stopped. If we continue through the
vectors in $A$, we will eventually end at a point,
\[ (\l_i e_i,\dots,\l_me_m,\l_1e_1,\dots,\l_{i-1}e_{i-1})\in \GL_m(\C), \]
which is diagonal, except the columns may have been
permuted if $n$ is not divisible by $m$. Choosing a fixed path in
$\GL_m(\C)$ from $(e_i,\dots,e_{i-1})$ to $\Id$, scaling the columns and
appending it to all the other paths then gives a path $ \g_A$ with
\[ \g_A(0) = \Id, \qquad \g_A(1) =
(\l_1e_1,\dots,\l_me_m). \]
If we consider this as a path in $\GLB{m}$ we
get a loop, since $\g_A(1)\in\B_m$ is a diagonal matrix. The
assignment
\[ f(A) = \g_A \in \Omega(\GLB{m}) \]
defines a continuous function from $Y_{m,n}$ to the loop space, as the
paths will be close at every point if the vectors we start from are
close. Note that
this function descends to the quotient space $X_{m,n}=Y_{m,n}/\T^n$, since
elements of $Y_{m,n}$ that are equivalent under the
$\mathrm{T}^n$-action gives
paths in $\GL_m(\C)$ that are equivalent under the $\mathrm{B}_m$-action.
The goal of this chapter is to prove the following theorem:
\begin{theorem}
The map $f : X_{m,n} \to \Omega$ induces a map
\[ f : X_{m,\infty} \to \Omega. \]
This is a homotopy equivalence.
\end{theorem}
Since the space $\Omega$ is relatively well-studied, this gives
information about $X_{m,\infty}$. In particular, it allows us to
calculate the homology and cohomology of this space.
We want to show that the stabilisation map
\begin{align*}
s : X_{m,n} &\to X_{m,n+1} \\
A = (a_1,\dots,a_n) &\mapsto \left(La_1,\dots,La_n,
\begin{pmatrix}
1 \\
\vdots\\
1
\end{pmatrix}
\right) = (LA,Le_1)
\end{align*}
does not affect the map $f$, or more precisely that $f$ induces a
well-defined map on $X_{m,\infty}$. To do this we will work with the
subsequence $X_{m,nm}$ of the sequence $X_{m,n}$, as the two limits
are isomorphic. This is a purely formal fact that follows from
stabilising $X_{m,r}$ to $X_{m,mn}$, which gives a map
\[ X_{m,\infty} \to \varinjlim_{n} X_{m,mn}. \]
There is a similar map defined by considering $X_{m,mn}$ as the space
$X_{m,r}$ for $r = mn$, which also respects stablisation, and hence
defines a map
\[ \varinjlim_n X_{m,mn} \to X_{m,\infty}. \]
The composition of these two maps in either direction is the identity,
so the two limits are isomorphic.
We will show that the diagram
\[ \xymatrix{ X_{m,nm} \ar[rr]^{s^m}\ar[dr]_f & & X_{m,(n+1)m}
\ar[dl]^f \\
& \Omega & } \]
commutes up to homotopy. This is done by showing that the map $s^m$ is
homotopic to the map $\widetilde s$ that inserts the identity on the
last $m$ columns of $A \in X_{m,nm}$,
\[ \widetilde s (a_1,\dots,a_{nm}) =
(a_1,\dots,a_{nm},e_1,\dots,e_m). \]
If we connect $\Id$ and $L^{-m}$ by a path $\g$ in the space of
invertible lower-triangular matrices, we get a homotopy
\[ s_t(A) = (\g(t)L^mA,
\g(t)L^me_1,\g(t)L^{m-1}e_1,\dots,\g(t)Le_1) \]
which goes from $s_0 = s^m$ to the map
\[ s_1 : A \mapsto (A,e_1,L^{-1}e_1,\dots,L^{-m+1}e_1). \]
The matrix $T = (e_1,L^{-1}e_1,\dots,L^{-m+1}e_1)$ is upper-triangular
with entries given by the formula:
\[ T_{ij} = (L^{-j+1}e_1)_i = (-1)^{i-1}\binom{j-1}{i-1} \quad
\text{when } i \leq j. \]
This is proven by considering negative powers of $L$ and showing the
identity:
\[ (L^{-j+1})_{ik} = (-1)^{i-k}\binom{j-1}{i-k} \quad \text{when } i
\geq k. \]
This is automatically true when $j$ is one, so assume it is true for
a given $j$. We will then show it for $j+1$. We are considering the
entry
\[ (L^{-(j+1)+1})_{ik} = (L^{-j})_{ik} = (L^{-1}\cdot L^{-j+1})_{ik} =
\sum_{l=1}^m (L^{-1})_{il} \cdot (L^{-j+1})_{lk}. \]
But if we use that $L^{-1}$ has entries
\[ (L^{-1})_{il} =
\begin{cases}
-1 & l = i-1, \\
1 & l = i, \\
0 & \text{otherwise},
\end{cases} \]
this simplifies to
\begin{align*}
(L^{-(j+1)+1})_{ik} &= (-1)\cdot(-1)^{i-1-k}\binom{j-1}{i-1-k}
+ (-1)^{i-k}\binom{j-1}{i-k} \\
&= (-1)^{i-k}\cdot \left( \binom{j-1}{i-k-1} +
\binom{j-1}{i-k} \right)\\
&= (-1)^{i-k} \binom{j}{i-k} \\
&= (-1)^{i-k} \binom{(j+1)-1}{i-k}.
\end{align*}
Since we are considering only the first column, we get
\[ T_{ij} = (L^{-j+1}e_1)_i = (L^{-j+1})_{i1} =
(-1)^{i-1}\binom{j-1}{i-1}. \]
As $T$ is an upper-triangular matrix, it can be homotoped to
the identity without affecting the linear independence of the
columns, and the maps $s^m$ and $\widetilde s$ are therefore
homotopic. The composition
$f\circ \widetilde s$ is the same as $f$, except that the loop
$f\circ\widetilde s$ is stationary at the identity for a while before
it terminates. Putting it all together shows the following:
\begin{lemma}
\label{lem:homotopycommute}
The diagram
\[ \xymatrix{ X_{m,nm} \ar[rr]^{s^m}\ar[dr]_{f} & & \ar[dl]^f
X_{m,(n+1)m} \\
& \Omega & } \]
commutes up to homotopy and $f$ induces a map:
\[ f : X_{m,\infty} \to \Omega. \]
\end{lemma}
\section{Homotopy equivalence}
To show that the map $f : X_{m,\infty} \to \Omega$
gives an isomorphism of homotopy groups, we will again follow
\cite{milnor}. For each element $U$ of $\SUT{m}$, we can find a
geodesically convex open ball centered on $U$. Since $\SUT{m}$
is compact, there is some $\e > 0$ such that any ball of radius $\e$
or less is geodesically convex. By the same argument, we can also
choose $\e$ so small that for any two points $A= (a_1,\dots, a_m)$ and
$B=(b_1,\dots,b_m)$ in $\SUT{m}$, with the distance from $A$ to $B$
less
than $\e$, any $m$ subsequent vectors in the sequence
$(a_1,\dots,a_m,b_1,\dots,b_m)$ are linearly
independent.
With this, we get an increasing sequence of open sets in $\Omega$,
defined by requiring the loops to be piecewise contained in a ball of
radius $\e$:
\[ \Omega_k = \set{\g \in \Omega \delim \Big\vert\delim
\g_{\left|\left[ \frac{j-1}{2^k},\frac{j}{2^k} \right]\right.} \text{ is
contained in an $\e$-ball $\mathrm{B}_\e$} }. \]
By the choice of $\e$ made above, we can define a map
\begin{align*}
\varphi_k : \Omega_k &\to X_{m,(2^k-1)m} \\
\g &\mapsto \left( \g\left(\frac{1}{2^k}\right),
\g\left(\frac{2}{2^k}\right), \dots,
\g\left(\frac{2^k-1}{2^k}\right) \right).
\end{align*}
The map is continuous since two paths being close in $\Omega_k$ means
that the maximum distance between them is small. But this translates
to the vectors of the image being close together. The maps $\varphi_k$
and $\varphi_{k+1}$ are related in the following fashion:
\begin{lemma}
\label{lem:com}
The following diagram commutes up to homotopy:
\[ \xymatrix{\Omega_k \ar@{^(->}[r] \ar[d]^{\varphi_k} & \Omega_{k+1}
\ar[d]^{\varphi_{k+1}} \\
X_{m,(2^k-1)m} \ar[r]^-{s} & X_{m,(2^{k+1}-1)m}. } \]
\end{lemma}
\begin{proof}
The proof is by direct computation.
\begin{align*}
\varphi_{k+1} &= \left( \g \mapsto \left(
\g\left(\frac{1}{2^{k+1}}\right),
\g\left(\frac{2}{2^{k+1}}\right), \dots,
\g\left(\frac{2^{k+1}-1}{2^{k+1}}\right) \right)\right) \\
&\simeq \left(\g \mapsto \left( \g\left(\frac{2}{2^{k+1}}\right),
\g\left(\frac{2}{2^{k+1}}\right),
\g\left(\frac{4}{2^{k+1}}\right), \dots,
\g\left(\frac{2^{k+1}}{2^{k+1}}\right) \right)\right) \\
&\simeq \left(\g\mapsto \left( \g\left(\frac{1}{2^{k}}\right),
\g\left(\frac{2}{2^{k}}\right), \dots,
\g\left(\frac{2^k-1}{2^k}\right), \Id,\dots,\Id \right) \right)\\
&= s\circ \varphi_k.
\end{align*}
The first homotopy is given by sliding along the minimal geodesic
between the points, which gives an allowed path by our choice of
$\e$. The second homotopy is somewhat similar: If $(a_1,\dots,a_m)$
and $(b_1,\dots,b_m)$ are two elements of $\SUT{m}$ and we have an
element of $X_{m,n}$ of the form
\[ \left( \dots, a_1,\dots,a_m,a_1,\dots,a_m,b_1,\dots,b_m,\dots
\right), \]
then we can change the middle columns one by one, as we did in
the construction of $f$, and get a path in $X_{m,n}$ that changes
the middle from $(a_1,\dots,a_m)$ to $(b_1,\dots,b_m)$:
\begin{align*}
\left(a_1,\dots,a_m,a_1,a_2,\dots,a_m,b_1,\dots,b_m \right)
&\leadsto
\left(a_1,\dots,a_m,b_1,a_2,\dots,a_m,b_1,\dots,b_m\right) \\
&\leadsto
\left(a_1,\dots,a_m,b_1,b_2,\dots,a_m,b_1,\dots,b_m\right) \\
&\leadsto
\left(a_1,\dots,a_m,b_1,b_2,\dots,b_m,b_1,\dots,b_m\right).
\end{align*}
This allows us to take all the repetitions
$\left(\dots,\g(t),\g(t),\dots\right)$ and move them to the end as
diagonal matrices, which results in the stabilisation map $s$.
\end{proof}
The above lemma shows that the maps $\varphi_k$ fit together to define
a map from $\Omega$ to $X_{m,\infty}$.
The idea is to use the maps $\varphi_k$ as homotopy inverses of the
loop space map $f$. We are now ready to prove the result.
\begin{theorem}
\label{thm:loekker}
The map $f : X_{m,\infty} \to \Omega$ is a homotopy
equivalence.
\end{theorem}
\begin{proof}
We will prove that $f$ is a weak homotopy equivalence, i.e. the map
\[f_* : \pi_i(X_{m,\infty}) \to \pi_i(\Omega)\]
is an isomorphism for all $i$. Since the
loop space is a CW complex and $X_{m,\infty}$ is a limit of CW
complexes, by respectively Corollary 17.2 and the appendix of
\cite{milnor}, this proves the theorem.
We will start with surjectivity. Let
\[ g : W \to \Omega \]
be a finite cell complex in $\Omega$. We want to show that we can
construct a map $\widetilde g$ from $W$ to $X_{m,\infty}$ such that
the composition $f\circ \widetilde g$ is homotopic to $g$.
The image of $g$ is compact and hence contained in $\Omega_k$ for
some $k$, since $\set{\Omega_k}_{k\in\N}$ is an open cover of
$\Omega$. Consider the map
\[ \widetilde g = \varphi_k\circ g : W \to X_{m,(2^k-1)m}. \]
Evaluating at a point $w\in W$ gives
\[ \widetilde g(w) = \varphi_k(\g_w) = \left(
\g_w\left(\frac{1}{2^k}\right),\dots,
\g_w\left(\frac{2^k-1}{2^k}\right) \right). \]
Evaluating $f$ on this gives a path connecting the points
$\g_w\left(\frac{j}{2^k}\right)$. But by our choice of $\e$, any
such path is homotopic to the path connecting these points with
minimal geodesics, so $f\circ\widetilde g \simeq g$.
To show injectivity, let
\[ g : W \to X_{m,\infty}\]
be a finite cell complex with $f\circ g$ null-homotopic. We want to
show that $g$ is null-homotopic in $X_{m,\infty}$. In the
direct limit, a compact set is contained in the image of $X_{m,n}$
for some $n$ (see \cite[Proposition~A.1]{hatcher} or
\cite[Chapter~9.4]{may}), so $g$ is represented by a map:
\[ g : W \to X_{m,n}. \]
We will work with this representative. The map $f\circ
g$ is null-homotopic in $\Omega$ and the image of the homotopy is
compact, so it is
contained in $\Omega_k$ for some $k$. By applying $\varphi_k$, we
get a function
\[ \varphi_k\circ f\circ g : W \to X_{m,(2^k-1)m}. \]
Since $f\circ g$ is null-homotopic, this function is
null-homotopic. But by the same argument that was used in
Lemma \ref{lem:com}, the map
\[ \varphi_k(f(g(w))) = \left(
\g_w\left(\frac{1}{2^k}\right),\dots,\g_w\left(\frac{2^k-1}{2^k}
\right)\right) \]
is homotopic to the stabilisation map
\[ s(g(w)) = \left[g(w),\Id,\dots,\Id\right], \]
with the homotopy given by sliding along the path $\g_w=f(g(w))$.
So the map $g$ can be stabilised to a null-homotopic map, so it must
be null-homotopic in $X_{m,\infty}$. This shows that $f_*$ is
injective on homotopy groups.
\end{proof}
Since $f$ is a homotopy equivalence, it follows that it induces
an isomorphism on homology and cohomology groups. The Pontrjagin
homology ring of the loop space, with product given by loop
concatenation, has been calculated in \cite[Theorem~4.1]{grbic} as
\begin{align*}
H_*(\Omega(\SUT{m})) \cong T(x_1,\dots,x_{m-1}) \otimes
\Z[y_1,\dots,y_{m-1}]/I,
\end{align*}
where $T(x_1,\dots,x_{m-1})$ is the tensor algebra generated by
$x_1,\dots,x_{m-1}$. The product has the relation $x_k^2 =
x_px_q+x_qx_p = 2y_1$ for $1\leq k,p,q\leq m-1$ and $p\neq q$. The
degree of $x_i$ is 1 and the degree of $y_i$
is $2i$. By the above result, this is also the homology of
$X_{m,\infty}$. The groups are finitely generated and torsion free, so
by the Universal Coefficient Theorem the homology and cohomology
groups are isomorphic as abelian groups. Note that the first
cohomology group,
\[ H^1(\SUT{m}) \cong \Z^{m-1}, \]
is the same as the first cohomology of $X_{m,n}$, as
computed in Corollary \ref{cor:1-kohom}.
\section{Other coefficients}
\label{sec:coef}
In the above, we did not use the structure of the complex numbers in
any noticeable way. In particular, we could replace them by the reals
or the quaternions without any change to the proofs. This immediately
gives us two theorems, very similar to Theorem
\ref{thm:loekker}. These require computing Borel subgroups and maximal
tori for various spaces, an overview of which can be found in
e.g. \cite[Example 6.7]{malle}.
\begin{theorem}
The space $X_{m,\infty}^\R$ is homotopy equivalent to the loop space
of the real flag manifold,
$\Omega(\GL_m(\R)/\mathrm{B}_m) = \Omega(\SO_m/\mathrm{T}^{k})$, where
$\SO_m$ are the $m \times m$ special orthogonal matrices and
$\mathrm{T^k}$ is a maximal torus in the special orthogonal group.
By \cite{grbic}, the Pontrjagin homology ring of this space depends
on whether $m$ is odd or even. For odd $m$, it is
\[ H_*(\Omega(2k+1)/\T^k) \cong T(x_1,\dots,x_k)\otimes
\Z[y_1,\dots,y_{k-1},2y_n,\dots,2y_{2k-1}] / I, \]
where $x_i$ is degree 1, the degree of $y_i$ and $2y_i$ is
$2i$, and $I$ is generated by the relations
\begin{align*}
x_1^2-y_1, x_i^2-x_{i+1}^2&\quad 1 \leq i \leq k-1, \\
x_a x_b + x_b x_a&\quad a \neq b, \\
y_i^2-2y_{i-1}y_{i+1}+\dots\pm 2y_{2i}&\quad 1 \leq i \leq k-1.
\end{align*}
For even $m$, the homology ring is
\[ H_*(\Omega(2k)/\T^k) \cong T(x_1,\dots,x_k)\otimes
\Z[y_1,\dots,y_{k-2},y_{k-1}+z,y_{k-1}-z,2y_k\dots,2y_{2k-2}] /
I, \]
where $x_i$ is degree 1 and anything involving a $y_i$ is of degree
$2i$. The relations are generated by
\begin{align*}
x_1^2-y_1, x_i^2-x_{i+1}^2& \quad 1 \leq i \leq k-1, \\
x_ax_b+x_bx_a& \quad a\neq b, \\
y_i^2-2y_{i-1}y_{i+1}+\dots\pm 2y_{2i}&\quad 1 \leq i \leq k-2,
\\
(y_{n-1}+z)(y_{n-1}-z)-2y_{n-2}y_n+\dots\pm 2y_{2n-2}&.
\end{align*}
\end{theorem}
\begin{theorem}
The space $X_{m,\infty}^\Hq$ is homotopy equivalent to the loop
space of the quaternionic flag manifold
$\Omega(\GL_m(\Hq)/\mathrm{B}_m) = \Omega(\mathrm{Sp}_m/\T^m)$.
Here, $\mathrm{Sp}_m$ is the symplectic group and $\T^m$ is a
maximal torus.
The homology ring is
\[ H_*(\Omega(\mathrm{Sp}_m/\T^m)) \cong T(x_1,\dots,x_m)\otimes
\Z[y_2,\dots y_m] / I, \]
with $x_i$ of degree 1 and $y_i$ of degree $4i-2$. The relations are
\begin{align*}
x_i^2 = x_j^2&,\\
x_a x_b = -x_b x_a& \quad a\neq b.
\end{align*}
\end{theorem}
This gives the homology of $X_{m,\infty}^\R$ and
$X_{m,\infty}^\Hq$.
\section{Other stabilisations}
\label{sec:stabilitet}
Now that we know the space $X_{m,\infty}$, we can consider different
ways of reaching this space from $X_{m,n}(\sigma)$. The
first thing to note is that the space $X_{m,\infty}$, so far described
as sequences of linearly independent vectors $(v_1,v_2,\dots,v_k)$ in
$\C^m$ of varying length, may as well be described as infinite
sequences $(v_i)_{i=1}^\infty$ for which there is some $k$ such that
the matrix $(v_{k+rm +1},v_{k+rm +2},\dots,v_{k+(r+1)m})$ is the
identity matrix for all integers $r \geq 0$, with the same
requirements for
linear independence. We could also consider sequences that are
infinite in the other direction, or
doubly infinite sequences
that start and end with the identity matrix,
and get the same space. Likewise, $X_{m,\infty}(\sigma)$ could be
defined as doubly infinite series that start with the identity matrix
and end with infinite copies of $\sigma$. The homeomorphisms between
these spaces are the obvious ones of inclusion or cutting off the tail
or start of an infinite sequence. As we saw in the proof of Lemma
\ref{lem:homotopycommute}, the two inclusions of $X_{m,n}$ in
$X_{m,\infty}$ by either
using the stabilisation map $s$ or by adding on the identity matrix
are homotopic, so this does not change any of the above results. To
differentiate the two ways of stabilising, we will refer to the
inclusion by adjoining an
infinite sequence of identity matrices on the left as
\[ L_n : X_{m,n}(\sigma) \to X_{m,\infty}(\sigma). \]
This map is important since it respects the
filtration defined in Section \ref{sec:filtration},
\[ F_0 \subset F_1 \subset \dots \subset X_{m,n}(\sigma), \]
and hence defines a map between the spectral sequences of
$X_{m,n}(\sigma)$ and $X_{m,\infty}(\sigma)$. These spectral sequences
have not played a role so far for $\sigma\neq \Id$, but are defined
exactly as for $X_{m,n}$.
We will be using $L_n$ to prove the following theorem.
\begin{theorem}
\label{thm:kollaps}
The spectral sequence for $X_{3,\infty}(\sigma)$ collapses on the
second page, with all cohomology concentrated in the first column.
\end{theorem}
At first this theorem may seem slightly strange, but we know
that the spaces $X_{3,\infty}(\sigma)$ are homotopy equivalent for all
$\sigma$, since they are all equivalent to the loop space
$\Omega(\SU_3/\T^2)$, so the theorem
just tells us that all the cohomology of $X_{3,\infty}(\sigma)$ can be
found in the $(X_{3,\infty})_{\emptyset}$-part of $F_0 \cong
(X_{3,\infty})_{\emptyset}\times T^2$.
\begin{proof}
We will show that the spectral sequence for $X_{3,\infty}(\sigma)$
is determined entirely by the torus-part. To see this, note that
since all the spaces $X_{3,\infty}(\sigma)$ are equivalent, they all
have the same cohomology ring $R$. The first page of the spectral
sequence has the form
\[ R \otimes H^*(\T^2) \to (R\oplus R)\otimes H^*(T\times (D,D-0))
\to R\otimes H^*(D_2,D_2-0). \]
The boundary map $d_1$ is an inclusion of spaces
followed by the boundary map for the pair $(D,D-0)$,
\[ d_1 = \delta \circ \imath^*, \]
and we would like to show that the inclusion map
essentially does nothing. We know that for cohomology classes $r$
and $t$, we have
\[ \imath^*(r \otimes t) = \imath^*((r\otimes 1)\cdot (1\otimes t))
= \imath^*(r\otimes 1) \imath^*(1\otimes t), \]
so we need to know that happens in both these cases. Instead of
reducing all the way to what would correspond to $X_{3,n-1}(\tau)$,
as we have often done in the finite-dimensional case, it is in this
case simpler to only use
the torus action to change all of
the non-zero entries in the last column to one and avoid the
multiplication with the lower-triangular matrix. The inclusion
can be illustrated by considering a concrete example. For the
case
\[\imath : X_{3,\infty}(\Id)_{\set{1}}\times(D-0) \to
X_{3,\infty}(\Id)_{\emptyset}, \]
we get the map
\[ \left(\begin{pmatrix}
\matrow{r_1} & 0 \\
\matrow{r_2} & t_2 \\
\matrow{r_3} & 1
\end{pmatrix},\varepsilon\right) =
\left(\begin{pmatrix}
\matrow{r_1} & 0 \\
\matrow{r_2} & 1 \\
\matrow{r_3} & 1
\end{pmatrix}, t_2, \varepsilon\right) \mapsto
\left(\begin{pmatrix}
\matrow{\frac{1}{\varepsilon}r_1} & 1 \\
\matrow{r_2} & 1 \\
\matrow{r_3} & 1
\end{pmatrix}, t_2,\varepsilon\right). \]
When $\varepsilon$ is a fixed, small real number, we get a homotopy
commutative diagram
\[ \xymatrix{ X_{3,\infty}(\Id)_{\set{1}} \ar[d]^{\imath}
\ar@{^(->}[r] & X_{3,\infty}(\Id) \\
X_{3,\infty}(\Id)_{\emptyset} \ar@{^(->}[ur] }, \]
where the homotopy is given by first multiplying the top row with a
number moving from one to $\varepsilon$ and then moving the last
coordinate of the top row to zero, illustrated below:
\begin{align*}
\begin{pmatrix}
\matrow{\frac{1}{\varepsilon}r_1} & 1 \\
\matrow{r_2} & 1 \\
\matrow{r_3} & 1
\end{pmatrix}
&\leadsto
\begin{pmatrix}
\matrow{r_1} & \varepsilon \\
\matrow{r_2} & 1 \\
\matrow{r_3} & 1
\end{pmatrix} \\
&\leadsto
\begin{pmatrix}
\matrow{r_1} & 0 \\
\matrow{r_2} & 1 \\
\matrow{r_3} & 1
\end{pmatrix}.
\end{align*}
But the inclusions into $X_{3,\infty}(\Id)$ are homotopy
equivalences, since they correspond to considering paths with a
different endpoint in the loop formulation and we know these are all
equivalent. Hence we must have $\imath$ a homotopy equivalence as
well, and so it is an isomorphism on cohomology groups:
\[ \imath^*(r\otimes 1) = r\otimes 1. \]
Likewise, we can see that if we choose our matrix to be e.g.
\[ A = \begin{pmatrix}
1 & 0 & 0 \\
0 & 1 & 1 \\
0 & 0 & 1
\end{pmatrix}, \]
the inclusion does not change the rows of $A$ since we have
quotiented out the torus action on the columns. Hence we conclude
that we preserve the other part of the cohomology as well. All in
all, we get
\[ d_1(r\otimes t) = r \otimes \delta(t). \]
But we know the map $\delta$, and hence we can compute the map on
the first
column as
\begin{align*}
d_1(r \otimes t_1t_2) &= ( r\otimes t_2, - r\otimes t_1) \otimes
[D,D-0], \\
d_1(r \otimes t_1) &= (0, -r\otimes t_1) \otimes [D,D-0], \\
d_1(r \otimes t_2) &= (r\otimes t_2,0) \otimes [D,D-0], \\
d_1(r \otimes 1) &= 0.
\end{align*}
By doing the exact same thing for the second column, leaving out
the class $[D,D-0]$ to ease on notation, we see
\begin{align*}
d_1( r_1\otimes t_2, r_2\otimes t_1) &= (r_1 + r_2) \otimes
[D_2,D_2-0], \\
d_1( r_1 \otimes t_2, 0) &= 0, \\
d_1( 0, r_2\otimes t_1) &= 0.
\end{align*}
But by using this we can compute the second page and see that it is
concentrated as $R\otimes 1$ in the first column, as desired.
\end{proof}
\section{Cohomological stability}
\label{sec:costa}
Since we now know the space $X_{m,\infty}(\sigma)$ and its cohomology, we
would like to relate it to the cohomology of
$X_{m,n}(\sigma)$, for finite $n$. We once again specialise to the
case $m = 3$ in
everything that
follows. First, we need an offset that depends on the permutation
we are considering.
Define a function $\lambda$ on the symmetric group of three elements
by
\begin{align*}
\lambda(\sigma) &=
\begin{cases}
1 & \sigma = \Id, \\
2 & \sigma \in \set{(1\,2),(2\,3),(1\,2\,3)},
\\
3 & \sigma \in \set{(1\,3\,2), (1\,3)}.
\end{cases}
\end{align*}
We will need to distinguish the spectral sequences for various
permutations. We will refer to the groups of the spectral sequence
used to compute the cohomology of $X_{3,n}(\sigma)$ as
$E_r^{p,q}(n,\sigma)$. We will also use the notation
$\sigma_{\emptyset}$ for the permutation that shows up in the
column $p=0$ when using the spectral sequence to compute cohomology,
corresponding to
\[X_{3,n}(\sigma)_{\emptyset} \cong
X_{3,n-1}(\sigma_{\emptyset})\times \T^2,\]
and likewise for $\sigma_{\set{1}},
\sigma_{\set{2}}$ and $\sigma_{\set{1,2}}$.
This section will prove the following theorem:
\begin{theorem}
\label{thm:kostab}
If $n \geq k + \lambda(\sigma)$, the stabilisation map
\[ L_{n}^* : H^k(X_{3,\infty}(\sigma)) \to H^k(X_{3,n}(\sigma)) \]
is an isomorphism.
\end{theorem}
\begin{proof}
The proof will proceed by induction on $k$. Our induction hypothesis
is divided into two parts, we require the stabilisation to be an
isomorphism when $n$ is greater than $k+\lambda(\sigma)$ and
injective when $n$ is $k+\lambda(\sigma)-1$. The exact statement we
want to prove is the following:
\begin{hypothesis}[$I_{k-1}$]
For all $r \leq k-1$ and all $\sigma \in S_3$, the following two
statements hold:
\begin{itemize}
\item The map $L_{n}^* : H^r(X_{3,\infty}(\sigma)) \to
H^r(X_{3,n}(\sigma))$ is an isomorphism for $n \geq r +
\lambda(\sigma)$.
\item The map $L_{n}^* : H^r(X_{3,\infty}(\sigma)) \to
H^r(X_{3,n}(\sigma))$ is injective for $n \geq r +\lambda(\sigma)
-1$.
\end{itemize}
\end{hypothesis}
By direct calculation, the
groups $H^0(X_{3,n}(\sigma))$ are all $\Z$ when $n \geq
\lambda(\sigma)-1$, so the isomorphism part of $I_0$ is
automatically true while the injective map also becomes an
isomorphism.
We now assume the hypothesis $I_{k-1}$ and want to prove $I_k$. Before
we start, we consider the spectral sequence for $X_{3,n+1}(\sigma)$
and observe how the different permutations that show up relate to each
other. This is summarised in the following table:
\[
\begin{array}{|c|cccc|}
\hline
\sigma & \multicolumn{1}{c|}{\sigma_\emptyset}
& \multicolumn{1}{c|}{\sigma_{\set{1}}}
& \multicolumn{1}{c|}{\sigma_{\set{2}}} & \sigma_{\set{1,2}} \\
\hline
\Id & \Id & (1\;2) & (2\;3) & (1\;3\;2) \\
(1\;2) & \Id & (1\;2\;3) & \Id & (1\;3) \\
(2\;3) & \Id & (1\;2) & \Id & (1\;2) \\
(1\;2\;3) & \Id & \Id & \Id & \Id \\
(1\;3\;2) & (2\;3) & (1\;2\;3) & (2\;3) & (1\;2\;3) \\
(1\;3) & (2\; 3) & (2\;3) & (2\;3) & (2\;3) \\
\hline
\end{array}
\]
By considering the table, we can see the following inequalities for
the function $\lambda$:
\begin{align*}
\lambda(\sigma) &\geq \lambda(\sigma_{\emptyset}), \\
\lambda(\sigma) &\geq \lambda(\sigma_{\set{1}}) - 1, \\
\lambda(\sigma) &\geq \lambda(\sigma_{\set{2}}) - 1, \\
\lambda(\sigma) &\geq \lambda(\sigma_{\set{1,2}}) - 2. \\
\end{align*}
These will be used during the proof.
\subsection{Surjectivity of $s^*$}
Let $\sigma$ be a permutation in $S_3$ and let $n \geq k +
\lambda(\sigma)$. We start by showing that the map
\[ s^* : H^k(X_{3,n+1}(\sigma)) \to
H^k(X_{3,n}(\sigma_{\emptyset})) \]
is surjective.
To see this, we will look at the spectral sequences for
$X_{3,\infty}(\sigma)$ and $X_{3,n+1}(\sigma)$, and the maps between
them induced by the stabilisation map $L^*_{n+1}$. This is
summarised in the following diagram.
\[ \xymatrix{ H^k(X_{3,\infty}(\sigma))
\ar@{^(->}[r]^-{\imath^*}\ar[d]^{L^*_{n+1}} & E_1^{0,k}(\infty,\sigma)
\ar[r]^{d_1}\ar[d]^{A_0} &
E_1^{1,k}(\infty,\sigma) \ar[r]^{d_1}\ar[d]^{A_1} &
E_1^{2,k}(\infty,\sigma) \ar[d]^{A_2} \\
H^k(X_{3,n+1}(\sigma)) \ar@{^(->}[r]^-{\imath^*} &
E_1^{0,k}(n+1,\sigma) \ar[r]^{d_1} & E_1^{1,k}(n+1,\sigma)
\ar[r]^{d_1} & E_1^{2,k}(n+1,\sigma).
} \]
The upper row is exact, since it is part of the spectral sequence for
$X_{3,\infty}(\sigma)$ and this collapses at the second page by
Theorem \ref{thm:kollaps}. The
bottom row is a chain complex, by considering the spectral sequence,
and is exact at
$E_1^{0,k}(n+1,\sigma)$. To see this, consider the induction
hypothesis applied to the spectral sequence for $X_{3,n+1}(\sigma)$,
illustrated in Figure \ref{fig:surjektiv}. Since we choose to
stabilise from the left,
as discussed in Section \ref{sec:stabilitet} above, $L_n$
becomes a map of spectral sequences and respects the stabilisation map
$s$. We use the notation \colorbox{isomorphism}{$E_1^{p,q}$} to show
that the map $L_n^*$ is an isomorphism on $E_1^{p,q}$ according to
$I_{k-1}$.
\begin{figure}[ht]
\[
\begin{array}{|c|ccc|}
\hline
&
&
&
\\
\multirow{6}{*}{$k$}
& \colorbox{isomorphism}{$H^{k-2}(X_{3,n}(\sigma_{\emptyset}))
\otimes t_1t_2 $}
& \colorbox{isomorphism}{ $H^{k-2}(X_{3,n}(\sigma_{\set{1}}))
\otimes t_2 $}
& \multirow{6}{*}{\colorbox{isomorphism}{$
H^{k-2}(X_{3,n}(\sigma_{\set{1,2}}))$}} \\
& \colorbox{isomorphism}{$H^{k-1}(X_{3,n}(\sigma_{\emptyset}))
\otimes t_1$}
& \colorbox{isomorphism}{$H^{k-2}(X_{3,n}(\sigma_{\set{2}}))
\otimes t_1$}
& \\
& \colorbox{isomorphism}{$H^{k-1}(X_{3,n}(\sigma_{\emptyset}))
\otimes t_2 $}
& \colorbox{isomorphism}{$H^{k-1}(X_{3,n}(\sigma_{\set{1}}))
\otimes 1$} &\\
& H^k(X_{3,n}(\sigma_{\emptyset})) \otimes 1
& \colorbox{isomorphism}{$H^{k-1}(X_{3,n}(\sigma_{\set{2}}))
\otimes 1$} &\\
& & &\\
\hline& & & \\
\multirow{6}{*}{$k-1$}
& \colorbox{isomorphism}{$H^{k-3}(X_{3,n}(\sigma_{\emptyset}))
\otimes t_1t_2$}
& \colorbox{isomorphism}{$H^{k-3}(X_{3,n}(\sigma_{\set{1}}))
\otimes t_2$}
& \multirow{6}{*}{\colorbox{isomorphism}{$
H^{k-2}(X_{3,n}(\sigma_{\set{1,2}}))$}} \\
& \colorbox{isomorphism}{$H^{k-2}(X_{3,n}(\sigma_{\emptyset}))
\otimes t_1$}
& \colorbox{isomorphism}{$H^{k-3}(X_{3,n}(\sigma_{\set{2}}))
\otimes t_1$} &\\
& \colorbox{isomorphism}{$H^{k-2}(X_{3,n}(\sigma_{\emptyset}))
\otimes t_2 $}
& \colorbox{isomorphism}{$H^{k-2}(X_{3,n}(\sigma_{\set{1}}))
\otimes 1$}
&\\
& \colorbox{isomorphism}{$H^{k-1}(X_{3,n}(\sigma_{\emptyset}))
\otimes 1$}
& \colorbox{isomorphism}{$H^{k-2}(X_{3,n}(\sigma_{\set{2}}))
\otimes 1$}
&\\
&& & \\
\hline
& & & \\
& \multirow{4}{*}{\dots}
& \multirow{4}{*}{\dots}
& \multirow{4}{*}{\dots}
\\
& & & \\
& & & \\
& & & \\
& & & \\
\hline && & \\
\multirow{6}{*}{$0$}
& \multirow{6}{*}{\colorbox{isomorphism}{$E_{1}^{0,0}$}}
& \multirow{6}{*}{\colorbox{isomorphism}{$E_1^{1,0}$}}
& \multirow{6}{*}{\colorbox{isomorphism}{$E_1^{2,0}$}} \\
& & &\\
& & &\\
& & &\\
& & &\\
\hline
\end{array}
\]
\caption{The $E_1$ page when we assume $I_{k-1}$ and $n \geq
k+\lambda(\sigma)$.}
\label{fig:surjektiv}
\end{figure}
If we pass to the second page of the spectral sequence, we
can see that at row $k$ the groups will no longer change. This is
because everything except $H^k(X_{3,n}(\sigma_{\emptyset}))\otimes 1$ is
isomorphic to the $X_{3,\infty}(\sigma)$-case, where the columns $p=1$
and $p=2$ disappear on the second page, according to Theorem
\ref{thm:kollaps}. We then know that
\[ H^k(X_{3,n+1}(\sigma)) = \ker d_1, \]
and the inclusion into $E_1^{0,k}(n+1,\sigma)$ is injective.
The map $A_1$, defined in the diagram, is an isomorphism, since it
is induced from the maps
\[ L^*_{n} : H^{k-1}(X_{3,\infty}(\sigma_{\set{i}})) \to
H^{k-1}(X_{3,n}(\sigma_{\set{i}})). \]
These are isomorphisms by applying the first part of $I_{k-1}$,
using the assumption
\[ n \geq k+\lambda(\sigma) \geq
(k-1)+\lambda(\sigma_{\set{i}}). \]
The map $A_2$ is an isomorphism by the exact same argument.
This allows us to do a diagram chase: Consider an element $x$ in
$E_1^{0,k}(n+1,\sigma)$. Then $d_1(x) = A_1(y)$ for a $y$ in
$E_1^{1,k}(\infty,\sigma)$. But
\[ A_2(d_1(y)) = d_1(A_1(y)) = d_1(d_1(x)) = 0, \]
so $y = d_1(z)$ for some $z$ in $E_1^{0,k}(\infty,\sigma)$ by
exactness. The element $x - A_0(z)$ is in the kernel of $d_1$,
\[ d_1(x - A_0(z)) = d_1(x) - A_1(d_1(z)) = d_1(x) - A_1(y) = 0. \]
So there is a $w$ in $H^k(X_{3,n+1}(\sigma))$ with $\imath^*(w)$
equal to $x-A_0(z)$. So we have found $w$ and $z$ with
\[ A_0(z) + \imath^*(w) = A_0(z) + x - A_0(z) = x, \]
showing that
\[ \xymatrix{ H^k(X_{3,n+1}(\sigma)) \oplus E_1^{0,k}(\infty,\sigma)
\ar[r]^-{\imath^* + A_0} & E_1^{0,k}(n+1,\sigma) } \]
is surjective. The map $s^*$ that we want to show is surjective is
the composition
\[ \xymatrix{ H^k(X_{3,n+1}(\sigma)) \ar[r]^-{\imath^*} &
E_1^{0,k}(n+1,\sigma) = H^k(X_{3,n}(\sigma_{\emptyset})\times T^2)
\ar[r]^-{\pr} & H^k(X_{3,n}(\sigma_{\emptyset})). } \]
The projection, which is induced from the inclusion of
$X_{3,n}(\sigma_{\emptyset})$ into the product, is surjective, so we
only need to show that the image of
$E_1^{0,k}(\infty,\sigma)$ under the map $\pr\circ A_0$ is contained
in the image of $s^*$ to
finish the proof of surjectivity of $s^*$. Consider the commutative
diagram
\[ \xymatrix{ E_1^{0,k}(\infty,\sigma) \ar[r]^{A_0} \ar[d]^{\pr}
& E_1^{0,k}(n+1,\sigma) \ar[d]^{\pr} \\
H^k(X_{3,\infty}(\sigma_{\emptyset})) \ar[r]^{L_n^*} &
H^k(X_{3,n}(\sigma_{\emptyset})). } \]
This shows that $\pr\circ A_0$ is contained in the image of
$L_n^*$. But the stabilisation map $L_n$ factors up to homotopy over
$s$, $L_n \simeq L_{n+1}\circ s$, so we get that
\[ \im(\pr\circ A_0) \subset \im L_n^* \subset \im s^*. \]
By applying the surjectivity of $\pr\circ (\imath^*+A_0)$ that was
just shown, we see that
\[ s^* : H^k(X_{3,n+1}(\sigma)) \to
H^k(X_{3,n}(\sigma_{\emptyset})) \]
is surjective.
\subsection{Injectivity}
We will now prove injectivity of $L_n^*$, so assume that $n \geq
k+\lambda(\sigma)-1$. As above, we will prove that
the stabilisation map
\[ s^* : H^k(X_{3,n+1}(\sigma)) \to
H^k(X_{3,n}(\sigma_{\emptyset})) \]
is injective. Consider the spectral sequence for computing
$H^k(X_{3,n+1}(\sigma))$. Figure \ref{fig:injektiv} is an image of
the first page of the spectral sequence, with
\colorbox{isomorphism}{$E_1^{p,q}$} denoting that $L_n^*$ is an
isomorphism on the group and \colorbox{injective}{$E_1^{p,q}$}
denoting that it is injective.
\begin{figure}[ht]
\[
\begin{array}{|c|ccc|}
\hline
& & & \\
\multirow{6}{*}{$k$}
& \colorbox{isomorphism}{$H^{k-2}(X_{3,n}(\sigma_{\emptyset}))
\otimes t_1t_2 $}
& \colorbox{isomorphism}{ $H^{k-2}(X_{3,n}(\sigma_{\set{1}}))
\otimes t_2 $}
& \multirow{6}{*}{\colorbox{injective}{$
H^{k-2}(X_{3,n}(\sigma_{\set{1,2}}))$}} \\
& \colorbox{isomorphism}{$H^{k-1}(X_{3,n}(\sigma_{\emptyset}))
\otimes t_1$}
& \colorbox{isomorphism}{$H^{k-2}(X_{3,n}(\sigma_{\set{2}}))
\otimes t_1$}
& \\
& \colorbox{isomorphism}{$H^{k-1}(X_{3,n}(\sigma_{\emptyset}))
\otimes t_2 $}
& \colorbox{injective}{$H^{k-1}(X_{3,n}(\sigma_{\set{1}}))
\otimes 1$} &\\
& H^k(X_{3,n}(\sigma_{\emptyset})) \otimes 1
& \colorbox{injective}{$H^{k-1}(X_{3,n}(\sigma_{\set{2}}))
\otimes 1$} &\\
& & &\\
\hline& & & \\
\multirow{6}{*}{$k-1$}
& \colorbox{isomorphism}{$H^{k-3}(X_{3,n}(\sigma_{\emptyset}))
\otimes t_1t_2$}
& \colorbox{isomorphism}{$H^{k-3}(X_{3,n}(\sigma_{\set{1}}))
\otimes t_2$}
& \multirow{6}{*}{\colorbox{isomorphism}{$
H^{k-2}(X_{3,n}(\sigma_{\set{1,2}}))$}} \\
& \colorbox{isomorphism}{$H^{k-2}(X_{3,n}(\sigma_{\emptyset}))
\otimes t_1$}
& \colorbox{isomorphism}{$H^{k-3}(X_{3,n}(\sigma_{\set{2}}))
\otimes t_1$} &\\
& \colorbox{isomorphism}{$H^{k-2}(X_{3,n}(\sigma_{\emptyset}))
\otimes t_2 $}
& \colorbox{isomorphism}{$H^{k-2}(X_{3,n}(\sigma_{\set{1}}))
\otimes 1$}
&\\
& \colorbox{isomorphism}{$H^{k-1}(X_{3,n}(\sigma_{\emptyset}))
\otimes 1$}
& \colorbox{isomorphism}{$H^{k-2}(X_{3,n}(\sigma_{\set{2}}))
\otimes 1$}
&\\
&& & \\
\hline
& & & \\
& \multirow{4}{*}{\dots}
& \multirow{4}{*}{\dots}
& \multirow{4}{*}{\dots}
\\
& & & \\
& & & \\
& & & \\
& & & \\
\hline && & \\
\multirow{6}{*}{$0$}
& \multirow{6}{*}{\colorbox{isomorphism}{$E_{1}^{0,0}$}}
& \multirow{6}{*}{\colorbox{isomorphism}{$E_1^{1,0}$}}
& \multirow{6}{*}{\colorbox{isomorphism}{$E_1^{2,0}$}} \\
& & &\\
& & &\\
& & &\\
& & &\\
\hline
\end{array}
\]
\caption{The $E_1$ page when we assume $I_{k-1}$ and $n \geq
k+\lambda(\sigma)-1$.}
\label{fig:injektiv}
\end{figure}
If the map
\[ s^* : H^k(X_{3,n+1}(\sigma)) \to
H^k(X_{3,n}(\sigma_{\emptyset})) \]
is not injective, there is an element $\alpha$ of the kernel of the
boundary map that is also in the group
\[ H^{k-1}(X_{3,n}(\sigma_{\emptyset})) \otimes t_1 \oplus
H^{k-1}(X_{3,n}(\sigma_{\emptyset})) \otimes t_2 \oplus
H^{k-2}(X_{3,n}(\sigma_{\emptyset})) \otimes t_1 t_2. \]
But on these last groups, the stabilisation map is an isomorphism,
so we get an element $\widehat{\alpha}$ in the corresponding group
in the spectral sequence for $X_{3,\infty}(\sigma)$. Since the
stabilisation maps are maps of spectral sequences, we have
\[ 0 = d_1(L_{n}(\widehat{\alpha})) = L_n(d_1(\widehat{\alpha})). \]
But the right hand side is the composition of $L_n$, which we know by
the induction hypothesis to be injective, and the boundary map in the
spectral sequence for $X_{3,\infty}(\sigma)$, which we know to be
injective on the groups we are considering by Theorem
\ref{thm:kollaps}. So $\widehat{\alpha}$ must
be zero, and hence so is $\alpha$. This shows that $s^*$ is injective
when $n \geq k + \lambda(\sigma) - 1$. Now we consider the different
choices of $\sigma$.
\subsubsection{$\sigma = \Id$}
Since $\sigma_{\emptyset}$ is $\Id$, and all the maps
\[ s^* : H^k(X_{3,j+1}) \to H^k(X_{3,j}), \quad j \geq n, \]
are surjective, we know that the cohomology
of $X_{3,\infty}$ is the inverse
limit of the cohomology groups of $X_{3,n}$ using the stabilisation
map, see e.g. \cite[Chapter 19.4]{may}.
This means
\[ H^k(X_{3,\infty}) = \set{ (a_i)\in \prod_{i=1}^\infty
H^k(X_{3,i}) \sdel s^*(a_{i+1}) = a_i \;\forall i}. \]
When considered in this way, the map $L_n^*$ is just the projection on
the corresponding factor. But if we project to $H^k(X_{3,n})$ and
get zero, we know that
the only possible lift to $H^k(X_{3,n+1})$ is zero, since $s^*$ is
injective. Continuing like this, we see that the only element of
$H^k(X_{3,\infty})$ that can project to zero is the zero element, and
hence $L_{n}^*$ is injective in this case.
\subsubsection{$\lambda(\sigma) \geq 2$}
If $\lambda(\sigma)=2$, we know $\sigma \in
\set{(1\;2),(2\;3),(1\;2\;3)}$,
$\sigma_{\emptyset} = \Id$ and that
\[ n \geq k + \lambda(\sigma) - 1 = k+1, \]
giving
\[ n-1\geq k = k +\lambda(\Id)-1. \]
Consider the following commutative diagram:
\[ \xymatrix{ H^k(X_{3,\infty}(\sigma)) \ar[r]^{s^*}_{\cong}
\ar[d]^{L_n^*} &
H^k(X_{3,\infty}) \ar[d]^{L_{n-1}^*} \\
H^k(X_{3,n}(\sigma)) \ar[r]^{s^*} & H^k(X_{3,n-1}).
} \]
By the previous case, the map on the right is injective and so the
composition $s^* \circ L_{n}^*$ is as well. Since the diagram
commutes, this means that the left vertical map, $L_n^*$ must be
injective, as desired.
The last case has $\sigma_{\emptyset} = (2\;3)$ and proceeds exactly
as the case $\lambda(\sigma) = 2$, using the injectivity that has
already been proved for this permutation and that $\lambda((2\;3))$ is
$\lambda(\sigma)-1$.
\subsection{Surjectivity}
We finish by proving that $L_n^*$ is surjective, so we assume $n \geq
k+\lambda(\sigma)$. We have already
worked with the stabilisation map $s^*$ and we now split into cases
based on $\sigma$.
\subsubsection{$\sigma = \Id$}
We still know that the cohomology group of $X_{3,\infty}$ is the
inverse limit of the
cohomology groups of $X_{3,n}$ and that the map $L_n^*$ is the
projection. But since we know all the maps
\[s^* : H^k(X_{3,j+1})\to H^k(X_{3,j}), \quad j \geq n, \]
are surjective, we can hit anything in $H^k(X_{3,n})$ by choosing a