-
Notifications
You must be signed in to change notification settings - Fork 7
/
low-dimensional-topology.tex
executable file
·4271 lines (3040 loc) · 446 KB
/
low-dimensional-topology.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
\input{"/Users/brandonwilliams/Documents/LaTeX Includes/notespreamble.tex"}
\input{"/Users/brandonwilliams/Documents/LaTeX Includes/extrapackages.tex"}
\input{"/Users/brandonwilliams/Documents/LaTeX Includes/extracommands.tex"}
\begin{document}
\title{\Large Low Dimensional Topology}
\author{Brandon Williams \\ \texttt{[email protected]}}
\maketitle
\tableofcontents
\newpage
\section{Transversality and Intersection Theory}
\label{Transversality and Intersection Theory}
\subsection{Immersions, Submersions and Transversality}
\label{Immersions, Submersions and Transversality}
Let $f : M \rightarrow N$ be a smooth map between smooth manifolds. A point $p \in M$ is said to be \textbf{regular} if $df_p : T_p M \rightarrow T_{f(p)} N$ is surjective, otherwise it is said to be \textbf{critical}. The image of a critical point is called a \textbf{critical value}, and a \textbf{regular value} in $N$ is any point that is not a critical value. We say that $f$ is an \textbf{immersion} at $p \in M$ if $df_p$ is injective, and a \textbf{submersion} at $p \in M$ if $df_p$ is surjective. Recall the inverse function theorem and its corollaries pertaining to immersions and submersions.
\begin{thm}[Inverse Function Theorem]
If $df_p : T_p M \rightarrow T_{f(p)} N$ is an isomorphism, then $f$ is a local diffeomorphism at $p$, i.e. there is a neighborhood $U$ of $p$ and $V$ of $f(p)$ such that $f|_U : U \rightarrow V$ is a diffeomorphism.
\end{thm}
\begin{cor}
If $f : M^m \rightarrow N^n$ is an immersion at $p \in M$, then there are coordinates $\phi : U \rightarrow \mathbb R^m$ around $p \in M$ and $\varphi : V \rightarrow \mathbb R^n$ around $f(p) \in N$ such that $\varphi \circ f \circ \phi^{-1}(x_1,\ldots,x_m) = (x_1,\ldots,x_m,0,\ldots,0)$.
\end{cor}
\begin{proof}
First let $\phi$ and $\varphi$ be any charts such that $\phi(p)=\varphi(p)=0$, and let $g = \varphi \circ f \circ \phi^{-1}$. Then $dg_0 : R^m \rightarrow R^n$ is injective since $f$ is an immersion, so we can change the coordinates so that
\[ dg_0 = \begin{pmatrix} I_m \\ 0 \end{pmatrix} \]
where $I_m$ is the $m \times m$ identity matrix and $0$ is the $(n-m) \times m$ zero matrix. We cannot apply the inverse function theorem to $g$ since $m < n$, but if we define $G : \phi(U) \times R^{n-m} \rightarrow R^n$ by $G(x,z) = g(x)+(0,z)$, then $dG_0 = I_n$. So $G$ is a local diffeomorphism, and $G \circ \varphi$ is another chart around $f(p)$ of $N$. In these coordinates we have that the local representation of $f$ is just $(x_1,\ldots,x_m) \mapsto (x_1,\ldots,x_m,0,\ldots,0)$.
\end{proof}
\begin{cor}
\label{local submersion theorem}
If $f : M^n \rightarrow N^m$ is an submersion at $p \in M$, then there are coordinates $\phi : U \rightarrow \mathbb R^n$ around $p \in M$ and $\varphi : V \rightarrow \mathbb R^m$ around $f(p) \in N$ such that $\varphi \circ f \circ \phi^{-1}(x_1,\ldots,x_n) = (x_1,\ldots,x_m)$.
\end{cor}
\begin{proof}
First let $\phi$ and $\varphi$ be any charts such that $\phi(p)=\varphi(p)=0$, and let $g = \varphi \circ f \circ \phi^{-1}$. Then $dg_0 : R^n \rightarrow R^m$ is injective since $f$ is a submersion, so we can change the coordinates so that
\[ dg_0 = \begin{pmatrix} I_m & 0 \end{pmatrix} \]
where $I_m$ is the $m \times m$ identity matrix and $0$ is the $m \times (n-m)$ zero matrix. We cannot apply the inverse function theorem to $g$ since $m < n$, but if we define $G : R^n \rightarrow \varphi(V) \times R^{n-m}$ by $G(x_1,\ldots,x_n) = (g(x),x_{m+1},\ldots,x_n)$, then $dG_0 = I_n$. So $G$ is a local diffeomorphism, and $G^{-1} \circ \phi$ is another chart around $p$ of $M$. In these coordinates we have that the local representation of $f$ is just $(x_1,\ldots,x_n) \mapsto (x_1,\ldots,x_m)$.
\end{proof}
\begin{cor}
\label{pre-image of regular value is submanifold}
If $q$ is a regular value of $f : M^n \rightarrow N^m$, then $f^{-1}(q)$ is a submanifold of $M$ with $\codim f^{-1}(q) = m$.
\end{cor}
\begin{proof}
Let $p \in f^{-1}(q)$, then $f$ is a submersion at $p$, so by \cref{local submersion theorem} we can find coordinates $\phi : U \rightarrow \mathbb R^n$ around $p$ and $\varphi : V \rightarrow \mathbb R^m$ around $q$ such that $\varphi \circ f \circ \phi^{-1}(x_1,\ldots,x_n)=(x_1,\ldots,x_m)$ and $\varphi(q)=(0,\ldots,0)$. Thus for $x \in U$ and $f(x)=q$ we have $\phi(x)=(0,\ldots,0,x_{m+1},\ldots,x_n)$ for some $x_{m+1},\ldots,x_n$. So $\phi$ restricted to $U \cap f^{-1}(q)$ provides a chart of the relatively open set $U \cap f^{-1}(q)$ into $\mathbb R^{n-m}$, making $f^{-1}(q)$ into a submanifold of dimension $n-m$.
\end{proof}
Consider the following simple application of \cref{pre-image of regular value is submanifold}. For real-valued functions $f_1,\ldots,f_k : M \rightarrow \mathbb R$, when is the set $Z$ of common zeros a manifold? If we set $f = (f_1,\ldots,f_k) : M \rightarrow \mathbb R^k$, then it is clear that $f^{-1}(0)$ is a manifold if $0$ is a regular value of $f$. Another way of looking at this is to see that $df_p = ((df_1)_p,\ldots,(df_k)_p) : T_p M \rightarrow T_{f(p)} \mathbb R^k$, and this is surjective if and only if $(df_1)_p,\ldots,(df_k)_p$ are linearly independent functionals. When this happens we say $f_1,\ldots,f_k$ are independent at $p$, and so we have the following corollary.
\begin{cor}
If $f_1,\ldots,f_k : M \rightarrow \mathbb R$ are independent at all points where they simultaneously vanish, then the set of common zeros $Z$ is a submanifold with $\dim Z = \dim M - k$.
\end{cor}
In the case of the above corollary we say that the submanifold $Z$ has been \emph{cut out} by the functions $f_1,\ldots,f_k$. However, not every submanifold can be globally cut out by independent functions.
Now that we know the pre-image of regular values are submanifolds, we naturally wonder if the pre-image of \emph{submanifolds} are submanifolds. Two smooth maps $f : M \rightarrow N$ and $g : L \rightarrow N$ are said to be \textbf{transversal} if at any point $p \in M$ and $q \in L$ for which $f(p)=g(q)$ we have
\[ df_{p}(T_p M) + dg_{q}(T_q N) = T_{f(p)} N \]
and we denote this by $f \pitchfork g$. Note that the above sum does \emph{not} have to be direct, and in fact most of the time will not be. If $g$ is just the inclusion of a submanifold $L$ of $N$, then we say $f$ is transverse to $L$ and write $f \pitchfork L$. If $f$ is also an inclusion of a submanifold $M$ of $N$, then we say the submanifolds $M$ and $L$ are transversal, and write $M \pitchfork L$.
\begin{prop}
\label{pre-image of submanifold under transversality}
If $f : M \rightarrow N$ is transverse to a submanifold $L$ in $N$, then $f^{-1}(L)$ is a submanifold of $M$ with $\codim f^{-1}(L) = \codim L$. In particular, if two submanifolds $M$ and $L$ of $N$ are transversal, then $M \cap L$ is another submanifold of $N$ and $\codim(M \cap L) = \codim M + \codim L$.
\end{prop}
An important feature of immersions, submersions, transversal intersections and embeddings is that they can be ``perturbed'' slightly, and they will remain in their respective class. To make this precise we need to define what we mean by perturbation. Fix a smooth map $f_0 : M \rightarrow N$, and let $F : M \times [-1,1] \rightarrow N$ be a smooth map such that $F(-,0) = f_0$. In fact, we will use the notation $f_t = F(-,t)$. Then any of the smooth maps $f_t : M \rightarrow N$ for $t \in [-1,1]$ are called \textbf{smooth perturbations} of $f_0$; that is, we just smoothly homotoped $f_0$ a little bit. We say that a statement $P$ defined on the space of smooth functions $C^\infty(M,N)$ is \textbf{stable} if for every $f \in C^\infty(M,N)$ such that $P(f)$ and every family of perturbations $F$ of $f$, then there is a small enough $\epsilon$ such that $P(f_t)$ for all $|t| < \epsilon$.
\begin{thm}
\label{stability theorem}
Let $M$ be compact. The following statements are stable as statements on $C^\infty(M,N)$.
\begin{enumerate}
\item $P(f) = f$ is an immersion.
\item $P(f) = f$ is an submersion.
\item $P(f) = f$ is a local diffeomorphism.
\item For a fixed submanifold $Z \subset N$, $P(f) = f$ is transversal to $Z$.
\item $P(f) = f$ is an embedding.
\item $P(f) = f$ is an diffeomorphism.
\end{enumerate}
\end{thm}
\begin{proof}
\sloppyspace
\begin{enumerate}
\item Suppose $f_0 : M^m \rightarrow N^n$ is an immersion, and take a family $F : M \times [-1,1] \rightarrow N$ perturbations of $f_0$, i.e. $F(-,0)=f_0$. The result will follow if we prove every point $(x,0) \in M \times [-1,1]$ has an open neighborhood $U_x$ such that $df_p$ is injective for all $p \in U$. For then we can take their union $U = \cup_x U_x$ and choose an $\epsilon>0$ small enough so that $M \times (-\epsilon,\epsilon) \subseteq U$. Since this is now just a local question we might as well assume $M$ is an open subset of $\mathbb R^m$ and $N$ is an open subset of $\mathbb R^n$. Since $(df_0)_x$ is injective, we have that the Jacobian matrix $(\partial (f_0)_i / \partial x_j(x))_{ij}$ contains at least $m$ linearly independent columns. We can of course assume that these columns are the first $m$ columns of the Jacobian. If we consider the Jacobian as a function of $(x,t) \in X \times [-1,1]$, then we see that the determinant of the first $m$ columns is a smooth function that is not equal to zero at $(0,x)$, hence we can find an $\epsilon>0$ small enough so that it is not zero for $(t,x)$, $|t|<\epsilon$. Therefore all the $f_t$'s are immersions for $|t|<\epsilon$.
\item The proof of this identical to 1 as now the Jacobian has $n$ linearly independent rows.
\item This follows directly from 1 since a local diffeomorphism is just an immersion for when $\dim M = \dim N$.
\item The condition of being transverse has an equivalent reformulation in terms of local submersions, so this follows from 2.
\item Recall that an embedding is a proper, injective immersion $f_0 : M \rightarrow N$. Since $M$ is compact, $f_0$ is automatically proper, so we only need to prove that if $f_0$ is injective, then $f_t$ is injective for $|t|$ less than some small $\epsilon>0$. We prove this by contradiction, so we can assume there is a sequence of numbers $t_i \to 0$ and distinct points $x_i \neq y_i$ such that $f_{t_i}(x_i) = f_{t_i}(y_i)$. Since $M$ is compact we can pass to a subsequence such that $t_i \to 0$, $x_i \to x$, and $y_i \to y$. If we define the function $G : M \times I \rightarrow N \times I$, then we have $G(x_i,t_i)=G(y_i,t_i)$ (by definition), hence
\[ G(x,0) = \lim_{i\to\infty} G(x_i,t_i) = \lim_{i\to\infty} G(y_i,t_i) = G(y,0) \]
Since $f_0$ is injective this implies $x=y$. The Jacobian of $G$ at $(x,0)$ is in block form
\[ \begin{pmatrix} & & & * \\ & d(f_0)_x & & \vdots \\ & & & * \\ 0 & \cdots & 0 & 1 \end{pmatrix} \]
But, $d(f_0)_x$ is injective, hence $dG_{(x,0)}$ has at least $m+1$ linearly independent columns and so $G$ is an immersion at $(x,0)$. This will remain true in a small neighborhood of $(x,0)$, which must contain the points $(x_i,t_i)$ and $(y_i,t_i)$ for large enough $i$. This contradicts that $G(x_i,t_i)=G(y_i,t_i)$, therefore no such sequence can exist, and so $f_t$ is injective for $|t|$ small enough.
\item
\end{enumerate}
\end{proof}
Now let us extend the ideas of transversality to manifolds with boundary. For a manifold $M$ with boundary, and a smooth map $f : M \rightarrow N$, the restriction of $f$ to the boundary $\partial M$ will be denoted by $\partial f$.
\unfinished
\subsection{Intersection Theory}
\label{Intersection Theory}
Recall the following classification of 1-dimensional manifolds and one of its immediate corollaries.
\begin{prop}
Every compact 1-dimensional manifold is homeomorphic to a disjoint union of circles and closed intervals.
\end{prop}
\begin{cor}
The number of boundary components of any compact 1-dimensional manifold is even.
\end{cor}
Consider a smooth map $f : M \rightarrow N$, with $M$ compact, and submanifold $Z \subseteq M$. By slightly perturbing $f$, if necessary, we can assume $f \pitchfork Z$ and so $f^{-1}(Z)$ is a submanifold of $M$. If $\codim Z = \dim M$, then $\codim f^{-1}(Z) = \codim Z = \dim M$, hence $f^{-1}(Z)$ is a 0-dimensional manifold. Since $M$ is compact, $f^{-1}(Z)$ consists of finitely many points, so we define the \textbf{mod-2 intersection number} of $f$ and $Z$ by $I_2(f,Z) = \#f^{-1}(Z) \modulo 2$. This is well-defined by the following proposition.
\begin{prop}
If $f,g : M \rightarrow N$ are homotopic, then $I_2(f,Z) = I_2(g,Z) \modulo 2$.
\end{prop}
\begin{proof}
Let $F : M \times I \rightarrow N$ be a homotopy between $f$ and $g$. We can perturb $F$ slightly, while keeping $F_0$ and $F_1$ fixed, to make $F \pitchfork Z$. Then $F^{-1}(Z)$ is a submanifold of $M \times I$ with $\codim F^{-1}(Z) = \codim Z = \dim M$, hence $\dim F^{-1}(Z)=1$. The boundary of $F^{-1}(Z)$ consists of
\[ F^{-1}(Z) \cap \partial (M \times I) = F^{-1}(Z) \cap (M \times \lcb 0 \rcb \cup M \times \lcb 1 \rcb) = f^{-1}(Z) \times \lcb 0 \rcb \cup g^{-1}(Z) \times \lcb 1 \rcb \]
Since $\# \partial F^{-1}(Z)$ is even, we must have that $\# f^{-1}(Z)$ and $\# g^{-1}(Z)$ are both even or both odd, hence $I_2(f,Z) = I_2(g,Z) \modulo 2$.
\end{proof}
We can define the mod-2 intersection number of two submanifolds $W,Z \subseteq M$ as $I_2(W,Z) := I_2(i,Z)$ where $i : W \hookrightarrow M$ is the inclusion. Although we can slightly perturb $W$ and $Z$ to create more intersection points (or remove intersection points), the above proposition says that points must be added in pairs.
\unfinished
\newpage
\section{Knot Theory}
\label{Knot Theory}
One of the reasons that low dimensional topology is so much harder to study than higher dimensions is the appearance of knots. In high dimensions one can unravel all knots to become ``trivial,'' but this does not happen in 3 dimensions. Fortunately we can get a combinatorial presentation of 3-manifolds from knots, but before that we have to cover some basics of knot theory.
\subsection{Basic Definitions}
\label{Basic Definitions}
Classically, a \textbf{knot} $k$ is just a smoothly embedded copy of $S^1$ into $S^3$. We could also embed our knots into $\mathbb R^3$, the theories would turn out to be equivalent, but $S^3$ is easier to think about since it is a nice compact 3-manifold. The complement of a knot in its surrounding manifold is called the \textbf{knot complement}, and its first homology group is quite simple.
\begin{prop}
\label{homology of knot complement in sphere}
If $k$ is a knot in $S^3$, then $H_1(S^3 \backslash k) = \mathbb Z$ and is generated by any meridian.
\end{prop}
\begin{proof}
Let $U$ be an open tubular neighborhood of $k$ in $S^3$, and let $V$ be the complement of the closure of a \emph{smaller} tubular neighborhood. Then $U$ is homotopy equivalent to $S^1$, $V$ is homotopy equivalent to the knot complement $S^3 \backslash k$, and $U \cap V$ is homotopy equivalent to a torus $S^1 \times S^1$. The Mayer-Vietoris sequences gives us the following exact sequence
\[ 0 \longrightarrow H_1(U \cap V) \stackrel{}{\longrightarrow} H_1(U) \oplus H_1(V) \stackrel{}{\longrightarrow} 0 \]
which is just
\[ 0 \longrightarrow \mathbb Z \oplus \mathbb Z \longrightarrow \mathbb Z \oplus H_1(S^3 \backslash k) \longrightarrow 0 \]
This implies that $H_1(S^3 \backslash k) = \mathbb Z$. Further, note that this map is of the form $(i_*,-j_*)$, where $i$ and $j$ are the inclusions of $U \cap V$ into $U$ and $V$, respectively. The group $H_1(U \cap V) = \mathbb Z \oplus \mathbb Z$ is generated by a meridian $[\mu]$ and longitude $[\ell]$, and we can choose this longitude such that $j_*[\ell]=0$. Since $[\ell]$ is mapped to an element generating the first factor of $\mathbb Z \oplus H_1(S^3 \backslash k)$, we have that $j_*[\mu]$ generates the second factor.
\end{proof}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%% TRY TO FINISH LATER %%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\comment{
\begin{prop}
\label{homology of knot complement in sphere}
If $k$ is a knot in $S^3$, then $H_n(S^3 \backslash k) = \mathbb Z$ for $n=0,1$ and $=0$ for $n \geq 2$.
\end{prop}
\begin{proof}
Since $S^3 \backslash k$ is a connected, non-compact 3-manifold, we already have $H_0(S^3 \backslash k) = \mathbb Z$ and $H_k(S^3 \backslash k) = 0$ for $k \geq 3$. Let $U$ be an open tubular neighborhood of $k$ in $S^3$, and let $V$ be the complement of the closure of a \emph{smaller} tubular neighborhood. Then $U$ is homotopy equivalent to $S^1$, $V$ is homotopy equivalent to the knot complement $S^3 \backslash k$, and $U \cap V$ is homotopy equivalent to a torus $S^1 \times S^1$. The Mayer-Vietoris sequence gives us the following exact sequences
\[ 0 \longrightarrow H_3(S^3) \stackrel{\Delta}{\longrightarrow} H_2(U \cap V) \stackrel{}{\longrightarrow} H_2(U) \oplus H_2(V) \stackrel{}{\longrightarrow} 0 \]
\[ 0 \longrightarrow H_1(U \cap V) \stackrel{}{\longrightarrow} H_1(U) \oplus H_1(V) \stackrel{}{\longrightarrow} 0 \]
The first exact sequence is really just
\[ 0 \longrightarrow \mathbb Z \stackrel{\Delta}{\longrightarrow} \mathbb Z \longrightarrow H_2(S^3 \backslash k) \longrightarrow 0 \]
To see that the connecting morphism $\Delta$ is an isomorphism recall how it is defined. Every homology class $\omega \in H_3(M)$ can be represented by a cycle of the form $c+d$, where $c$ and $d$ are cycles in $U$ and $V$ respectively. Since $\partial c = -\partial d$ we see that the image of $\partial c$ is contained in $U \cap V$. So we define $\Delta \omega = [\partial c] \in H_2(U \cap V)$. \unfinished
Whereas the second exact sequence is really just
\[ 0 \longrightarrow \mathbb Z \oplus \mathbb Z \longrightarrow \mathbb Z \oplus H_1(S^3 \backslash k) \longrightarrow 0 \]
Therefore $\mathbb Z^2 \cong \mathbb Z \oplus H_1(S^3 \backslash K)$, hence $H_1(S^3 \backslash k) = \mathbb Z$.
\end{proof}
}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%% /TRY TO FINISH LATER %%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
Given two knots $k_1,k_2$ in a manifold $S^3$, we say that they are \textbf{equivalent}, or \textbf{isotopic}, if they are ambiently isotopic. This means that there is a smooth 1-parameter family of maps $\varphi_t : S^3 \rightarrow S^3$ such that $\varphi_0 = \id$, $\varphi_1(k_1)=k_2$, and $\varphi_t$ is a diffeomorphism for all time $t$. A knot that is isotopic to the standardly embedded $S^1$ in $S^3$ is called the \textbf{unknot} or the \textbf{trivial knot}.
There is a nice way to visualize a knot $k$ in $S^3$. First, move the knot a little if needed so that it does not pass through the point at infinity. Now $k$ is embedded in $\mathbb R^3$. For a generic plane $\Pi$ in $\mathbb R^3$ not intersecting $k$, the projection $D$ of $k$ onto $\Pi$ is a closed curve with finitely many double points, and no other types of singularities. We can decorate these double points as an over or under crossing depending on whether one strand is above another relative to the plane. Such a projection $D$ is called a \textbf{knot diagram} of $k$. Conversely, given a closed curve $D$ in the plane which is embedded on the complement of finitely many double points and whose double points are decorated with under or over crossings, we can find a knot $k$ such that $D$ is a projection of $k$.
Clearly a single knot $k$ can admit many different knot diagrams, and so we hope there is a way to determine when two diagrams correspond to the same isotopy class of knot. It turns out there are 3 moves one can apply to a knot diagram, called the \textbf{Reidemeister moves}, such that two knot diagrams $D_1$ and $D_2$ are the diagrams of two isotopic knots if and only if one can get from one diagram to the other by the Reidemeister moves (see \cref{reidemeister moves}). We will use $R_1,R_2$ and $R_3$ to denote these moves. This reduces a lot of knot theory to combinatorial pictures, from which we can define a lot of knot invariants.
\begin{figure}[tb]
\centering
\includegraphics[scale=1]{graphics/reidemeister-moves}
\caption{Reidemeister Moves}
\label{reidemeister moves}
\end{figure}
A slight generalization of a knot is a link. An \textbf{$n$-component link} is a smooth embedding of $n$ copies of $S^1$. All of the above definitions and discussion has a direct generalization to links. Clearly if two links $L_1$ and $L_2$ are isotopic, then their number of components must be equal.
One can put an orientation on a knot $k$ by simply choosing a non-vanishing tangent vector field on $k$, or equivalently placing an arrow on any knot diagram of $k$. A knot has precisely 2 orientations and an $n$-component link has $2^n$ orientations. Given a knot diagram $D$ of an oriented knot $k$, we can define the \textbf{sign} of each crossing of $D$. We adopt the right hand rule for sign conventions, which means signs are determined as in \cref{crossing signs}. Note that the sign of a crossing is information we get once we have chosen a knot diagram, not from just a knot. The sum of all the signs of the crossing of a knot diagram $D$ is called the \textbf{writhe} of the diagram, and is denoted by $w(D)$. The writhe of a knot diagram is \emph{not} a knot invariant because it is not invariant under $R_1$. It is, however, a ribbon invariant, but we will not discuss that here.
\begin{figure}[tb]
\centering
\includegraphics[scale=1]{graphics/crossing-signs}
\caption{Crossing Signs}
\label{crossing signs}
\end{figure}
Let $k$ be a knot in $S^3$ and $D$ a diagram of $k$. A \textbf{resolution of a crossing}, or a \textbf{smoothing}, is the process of removing the double point and reconnecting the strands so that there is no crossing. There are two ways of doing this, called the 0-resolution and 1-resolution, shown in \cref{crossing resolution}. Each time one resolves a crossing the number of components of the diagram either stays the same or increases by one. A \textbf{resolution of a diagram} is the diagram resulting from resolving all crossings, which is just a finite collection of disjoint circles. Clearly if there are $n$ crossings in the diagram $D$, then there are $2^n$ resolutions, which are parameterized by $\lcb 0,1 \rcb^n$. If the knot $k$ has an orientation, then there is a canonical resolution of $D$, called the \textbf{oriented resolution}, where we resolve all the crossings of sign $+1$ by a 1-resolution and all crossings of sign $-1$ by a 0-resolution.
\begin{figure}[tb]
\centering
\[
\xy0;/r0.4pc/:
(-6,6)*{}="tl";
(6,6)*{}="tr";
(-6,-6)*{}="bl";
(6,-6)*{}="br";
"tl";"tr" **\crv{(0,1)};
"bl";"br" **\crv{(0,-1)};
(0,-10)*{\text{0-resolution}};
\endxy
\qquad \qquad
\xy0;/r0.4pc/:
(6,6)*{}="tl";
(-6,6)*{}="tr";
(6,-6)*{}="bl";
(-6,-6)*{}="br";
{\ar@{-}|{\hole \; \hole \; \hole \; \hole \; \hole \; \hole } "bl";"tr"};
{\ar@{-} "br";"tl"};
(0,-10)*{\text{crossing}};
\endxy
\qquad \qquad
\xy0;/r0.4pc/:
(-6,6)*{}="tl";
(6,6)*{}="tr";
(-6,-6)*{}="bl";
(6,-6)*{}="br";
"bl";"tl" **\crv{(-1,0)};
"br";"tr" **\crv{(1,0)};
(0,-10)*{\text{1-resolution}};
\endxy
\]
\caption{Ways to resolve a crossing}
\label{crossing resolution}
\end{figure}
A \textbf{Seifert surface} for a link $L$ in $S^3$ is a smoothly embedded, compact, connected and orientable surface $F$ in $S^3$ with boundary such that $\partial F = L$.
\begin{prop}[Seifert's Algorithm]
\label{Seifert's algorithm}
Every link $L$ in $S^3$ has a Seifert surface.
\end{prop}
\begin{proof}
We can construct the surface explicitly. Choose any orientation on $L$ and let $D$ be a diagram for $L$. Let $D'$ denote the the oriented resolution of $D$. Fill each component of $D'$ so that there is a disc for each circle, and connect the discs a twisting band for each crossing. The result is a surface with boundary equal to exactly $L$.
\end{proof}
Any link has many Seifert surfaces, and Seifert's algorithm almost never produces the surface of minimal genus. We define the \textbf{genus} of a link $L$ to be the minimum of genera of Seifert surfaces of $L$, and is denoted by $g(L)$. The \textbf{four-ball genus} of $L$ is the minimum of genera of compact, connected, oriented and smoothly embedded surfaces in $D^4$ whose boundary is $L \subset \partial D^4 = S^3$, and is denoted by $g^*(L)$.
\begin{lem}
\label{additivity of genus}
Let $k_1,k_2$ be two knots embedded in $S^3$. Then the genus satisfies $g(k_1 \# k_2) = g(k_1) + g(k_2)$.
\end{lem}
\begin{proof}
We can think of $k_1$ and $k_2$ embedded in the same sphere, but situated far apart. Let $F_1$ and $F_2$ be Seifert surfaces of minimal genus for $k_1$ and $k_2$, respectively. Then the boundary sum $F_1 \sharp F_2$ is a Seifert surface of $k_1 \# k_2$, and so we have proved
\[ g(k_1 \# k_2) \leq g(k_1) + g(k_2) \]
To prove the other inequality we fix a Seifert surface $F$ of minimal genus for $k_1 \# k_2$. We can fix a 2-sphere $\Sigma$ in $S^3$ so that $\Sigma \cap k_1 \# k_2$ consists of two points. Then, if the two arcs of $k_1 \# k_2$ between these two points are labelled $\alpha_1$ and $\alpha_2$, and we take a small arc $\beta$ in $\Sigma$ between these two points, then the curve $\alpha_1 + \beta$ is $k_1$ and the curve $\alpha_2 + \beta$ is $k_2$. By perturbing $\Sigma$ slightly we can guarantee that $\Sigma \cap F$ is just a collection of closed curves, as well as the curve $\beta$. We will perform surgery on these closed curves one by one to remove these intersections, so that the only intersection points left is the arc $\beta$.
Let $C$ be one of these closed curves with the additional property that the disc it bounds in $\Sigma$ does not contain any of the other closed curves. These ensures that if we push the interior of this disc a little off of $\Sigma$ (in either direction) it does not intersect $F$. Let $\hat F$ be the surface obtained from $F$ by cutting out an open annulus around $C$ (so now there are two more boundary components) and gluing in two discs, each parallel to the disc in $\Sigma$ bounded by $C$, and on either side of $\Sigma$. The result is another Seifert surface $\hat F$ of $k_1 \# k_2$. If the curve $C$ does not separate $F$, then $\hat F$ will have genus one less that $F$ (we just cut open a handle!), which is not possible since $F$ is assumed to have minimal genus. So, we must have that $C$ separates $F$, hence $\hat F$ consists of a disjoint union of a 2-sphere and a Seifert surface of $k_1 \# k_2$ that has one less intersection with $\Sigma$. Repeating this for all the closed curves in $\Sigma \cap F$ shows that we can assume that $\Sigma$ and $F$ intersect only at the arc $\beta$.
Now we can take $F_1$ to be the closure of the part of $F$ outside of the sphere $\Sigma$, and $F_2$ the closure of the part of $F$ inside the sphere. Then $F_1$ and $F_2$ are Seifert surfaces of $k_1$ and $k_2$ respectively, hence
\[ g(k_1) + g(k_2) \leq g(k_1 \# k_2) \]
\end{proof}
Given disjoint, oriented knots $k_1$ and $k_2$, let $D$ be a common projection. We can assign signs to the double points of $D$ just as before, except now the two strands might belong to different knots. The \textbf{linking number} $\lk(k_1,k_2)$ of $k_1$ and $k_2$ is defined to be the sum of the signs of all crossings where $k_1$ passes under $k_2$. In order to show that this definition is well-defined we need to see that it does not depend on the diagram chosen for $k_1 \cup k_2$. It is an easy matter to show that $\lk(k_1,k_2)$ is invariant under the Reidemeister moves, hence $\lk(k_1,k_2)$ is an invariant of the isotopy class of $k_1 \cup k_2$. It is also easy to check that $\lk(k_1,k_2)=\lk(k_2,k_1)$ and $\lk(-k_1,k_2) = -\lk(k_1,k_2)$, where $-k_1$ is the knot $k_1$ with opposite orientation.
There are two other ways of describing the linking number that will be useful later. Let $L = k_1 \cup k_2$ be the link with components our knots, and let $F$ be a Seifert surface for $L$. At each point on the boundary $k_1$ in $F$ we have a tangent vector $\tau$ that determines the orientation of $k_1$ and a vector $\nu$ pointing into the surface. Let $n$ be a vector normal to $F$ such that the ordered basis $(\tau,\nu,n)$ matches with the orientation of $S^3$. The normal vector $n$ can be extended to the rest of the surface $F$, and it determines an orientation. We can perturb $k_2$ slightly if needed so that $k_2$ intersects $L$ transversely in finitely many points. If we assign these points the sign $+1$ when $k_2$'s tangent vector points in the same direct as $n$ and $-1$ when it points in the opposite direction, then $\lk(k_1,k_2)$ is equal to the sum of these numbers.
The second way is to notice that the first homology of the knot complement $H_1(M\backslash k_1) = \mathbb Z$ is generated by the homology class $[\mu]$ of any meridian $\mu$ of $k_1$. The other knot $k_2$ determines a homology class $[k_2] \in H_1(M\backslash k_1)$, which must be an integer multiple of $[\mu]$. The linking number $\lk(k_1,k_2)$ is precisely this number, i.e. $[k_2] = \lk(k_1,k_2) \cdot [\mu]$.
A \textbf{framing} for a knot $k$ in $S^3$ is a choice of a normal vector field along $k$. If $X$ is a framing of $k$, then we can push $k$ a little bit in the direction of $X$ to get a new knot $k'$, and the integer $\lk(k,k')$ is called the \textbf{framing number} or \textbf{framing coefficient} of $X$. Conversely, given an integer $n$ we can find a parallel $k'$ of $k$ such that $\lk(k,k')=n$, and this curve is unique up to isotopy in $S^3 \backslash k$. So, we can think of framings as either normal vector fields, parallel knots, or just integers assigned to knots. A knot with a framing is called a framed knot, and a framed link is defined similarly. For an oriented, framed link $L$ with components $L_1,\ldots,L_n$ we define its \textbf{linking matrix} $[L]$ to be the symmetric matrix with $(i,j)$-entry equal to $\lk(L_i,L_j)$, where we use the convention that $\lk(L_i,L_i)$ is just the framing of the $i$-th component of $L$.
Given a knot $k$ in $S^3$ the meridian $\mu$ of $\partial N(k)$ chosen above is called the \textbf{canonical meridian}, and is always defined. However, a longitude $\ell$ of $\partial N(k)$ is not so clear how to define since it can wrap around $k$ in complicated ways. The ways in which $\ell$ can wrap around $k$ are parameterized by linking numbers, so we define the \textbf{canonical longitude} to be the curve $\ell$ in $\partial N(k)$ such that $\lk(k,\ell)=0$. This curve is unique up to isotopy, and will be useful in our discussion on knot surgeries. Given a diagram $D$ of a knot $k$ one can draw the canonical longitude as a parallel of $k$ that wraps $-w(D)$ full times around $k$, where $w(D)$ is the writhe of the diagram. Take note that we are using the linking number to define the canonical longitude even though our knot need not be oriented. However, we can given $k$ any orientation and point $\ell$ in the same direction, and then the linking number will not depend on the orientation we gave $k$.
\begin{figure}[tb]
\centering
\includegraphics[scale=.5]{graphics/canonical-longitude}
\caption{Examples of the canonical longitude and a non-canonical longitude}
\label{canonical longitude}
\end{figure}
In \cref{canonical longitude} we have three knots drawn (the bolded knots) with longitudes (not bolded). The first two longitudes are canonical, as one can easily calculate their linking numbers to be 0. However the third longitude, which seems like a more obvious choice for a longitude than the second, is \emph{not} canonical since its linking number is -3.
The fundamental group of the knot complement $\pi_1(S^3 \backslash k)$ is called the \textbf{knot group}, and clearly is an invariant of the knot. There is a simple algorithm for computing a presentation of $\pi_1(S^3 \backslash k)$ given a diagram of $D$. For a diagram $D$ of $k$, let $n+1$ be the number of crossings. Let $\alpha_0,\ldots,\alpha_n$ be the arcs in $D$ that start at a crossing, travel along the under-strand, and end at the next crossing that the strand goes under such that $\alpha_i$ starts at where $\alpha_{i-1}$ ends (we always take the subscripts of the $\alpha$'s to be mod $n$). Also orient these arcs such that $\alpha_i$ points towards the starting point of $\alpha_{i+1}$. For convenience, suppose that $k$ does not pass through $(0,0,1)$ (the point at infinity in $\mathbb R^3$), and let us use this as the base point in $S^3 \backslash k$. In terms of the knot diagram, we can think of the base point as sitting right where we (the viewers) are, and so loops based at $\infty$ leave from us, travel through the knot diagram, and then return to us. Hence, a loop in the knot diagram can be represented by just a curve segment (\emph{not} a loop), and we just assume the ends are connected to the point at infinity in the obvious way.
Generators of $\pi_1(S^3 \backslash k)$ are just short arc segments $x_0,\ldots,x_n$ in the diagram such that $x_i$ passes under $\alpha_i$, and oriented so that it crosses under $\alpha_i$ from the right to the left. Each crossing of the knot diagram gives a relation $r_i$ on the generators of the fundamental group. Suppose the crossing is where the arcs $\alpha_i,\alpha_j$ and $\alpha_{j+1}$ come together for some indices $i$ and $j$ (hence $\alpha_i$ is the over-strand). If the sign of the crossing is $+1$, then we add the relation $x_i x_j = x_{j+1} x_i$, and if the sign is $-1$, then we add the relation $x_j x_i = x_i x_{j+1}$. Note that we are computing the sign of these crossings with the orientation on $k$ induced by the orientations on the $\alpha$ arcs, which were chosen for convenience. If we changed those orientations we would get an equivalent presentation of the fundamental group. Let $r_i$ denote the relation obtained at the crossing where $\alpha_i$ and $\alpha_{i+1}$ meet.
\begin{prop}
The fundamental group of $S^3 \backslash k$ is isomorphic to $\< x_0,\ldots,x_n \st r_0,\ldots,r_n \>$.
\end{prop}
\begin{proof}
First we isotope $k$ so that it lies in $\lcb z = 0 \rcb \subset \mathbb R^3$, except at a crossing where the understand dips to the $\lcb z=-\epsilon \rcb$ plane for a short segment, which we call $\beta_i$ for the crossing corresponding to where $\alpha_i$ and $\alpha_{i+1}$ meet. We will decompose $S^3 \backslash k$ into a bunch of pieces and apply van Kampen's theorem. Let $A = \lcb z \geq -\epsilon \rcb \backslash k$, and let $B_i$ ($i=0,\ldots,n$) be a small open rectangular box around the segment $\beta_i$, but with $\beta_i$ remove and with a thin tube connected the box to $\infty$. Finally, let $C$ be the complement of the closure of $A \cup B_0 \cup \cdots \cup B_n$.
First we claim that $\pi_1(A)$ is a free group generated by $x_0,\ldots,x_n$. To see this let $\overline{\alpha}_i$ be the arc $\alpha_i$ whose endpoints dive straight down to connect to the plane $\lcb z=-\epsilon \rcb$, and let $\overline{k}$ be the union of these arcs. Then $\pi_1(S^3\backslash k)$ is isomorphic to $\pi_1(S^3 \backslash \overline{k})$, so we compute the later group. If we take a curtain below the arc $\overline{\alpha}_i$ to the plane $\lcb z=-\epsilon \rcb$ and then thicken it we get a ball $B_i$. Removing the $B_i$'s from $\lcb z \geq -\epsilon \rcb$ yields a simply connected space, so we determine what happens to the fundamental group when we add the pieces $B_i \backslash \overline{\alpha}_i$ back. The piece $B_i \backslash \overline{\alpha}_i$ is just a punctured disc, so its fundamental group is generated by a loop going around this hole once. Also the piece $B_i \backslash \overline{\alpha}_i$ intersects the space where we removed all the $B_j$'s from $\lcb z \geq -\epsilon \rcb$ is just a disc, hence we added one generator by adding the piece $B_i \backslash \overline{\alpha}_i$. Continuing this we get the claim.
\unfinished
\end{proof}
This presentation of the knot group is called the \textbf{Wirtinger presentation}.
\begin{example}
Consider the diagram of the trefoil knot $k$ shown in \cref{knot-group-trefoil} (we label our arcs by $A,B,C$ and generators of $\pi_1$ by $a,b,c$). The Wirtinger representation of the knot group is
\[ \pi_1(S^3 \backslash k) = \< a,b,c \st ba=ac, ac=cb, cb=ba \> \]
We can eliminate a generator from this presentation by seeing that $c = bab^{-1}$, hence another presentation of the fundamental group is given by
\[ \pi_1(S^3 \backslash k) = \< a,b \st bab=aba \> \]
If we set $x = aba$ and $y=ba$, then we see that $x^2 = abaaba = bababa = y^3$, hence yet another presentation is given by
\[ \pi_1(S^3 \backslash k) = \< x,y \st x^2 = y^3 \> \]
\begin{figure}[tb]
\centering
\includegraphics[scale=.5]{graphics/knot-group-trefoil}
\caption{Computing the knot group of the trefoil}
\label{knot-group-trefoil}
\end{figure}
\end{example}
\begin{example}
Let us compute the knot group of the $(p,q)$ torus knot $T_{p,q}$ (with $p,q \geq 1$ relatively prime). Arrange the knot to be on the boundary of the standardly embedded, unknotted solid torus $H$ in $S^3$. Then we can write $S^3 \backslash T_{p,q}$ as a union of $H \backslash T_{p,q}$ and $\overline{H^c} \backslash T_{p,q}$, each of which is clearly homotopy equivalent to a solid torus. The intersection of these sets is an annulus, and the fundamental group of each is freely generated by loops $x$ and $y$. By the Seifert-van Kampen theorem we have $\pi_1(S^3 \backslash T_{p,q})$ is generated by $x$ and $y$ with the relation $x^p = y^q$, hence
\[ \pi_1(S^3 \backslash T_{p,q}) = \< x,y \st x^p = y^q \> \]
\end{example}
\subsection{The Alexander Polynomial}
\label{The Alexander Polynomial}
We will now construct some classical invariants of knots and links in $S^3$. With some work these invariants can be extended to links in integral homology spheres, as we will see later. These combinatorial invariants have very interesting relationships with ``high-powered'' invariants, like Seiberg-Witten invariants and Heegaard Floer homology.
Fix a diagram $D$ of a knot $L$ in $S^3$ with $n$ components, and give the link any orientation. Suppose this diagram has $k$ double points and the oriented resolution of this diagram has $c$ connected components. Let us compute the Euler characteristic and genus of the Seifert surface $F$ constructed from Seifert's algorithm on $D$ (see \cref{Seifert's algorithm}). The Euler characteristic of $c$ disjoint discs is $c$, and attaching a band to these discs decreases the Euler characteristic by one. Therefore $\chi(F) = c - k$. On the other hand, an orientable surface of genus $g$ with $n$ boundary components has Euler characteristic $2-2g-n$. Setting these two Euler characteristics equal to each other and solving for $g$ shows that the genus of $F$ is
\begin{equation}
\label{genus of surface from Seifert's algorithm}
g = 1 + \frac{k-c-n}{2}
\end{equation}
One can compute that the first homology of a surface of genus $g$ with $n \geq 1$ boundary components is the free abelian group of rank $2g+n-1$ (just look at the handlebody decomposition of such a surface), therefore, by \eqref{genus of surface from Seifert's algorithm}, $H_1(F)$ is free abelian with rank $1+k-c$.
Now, for a general Seifert surface $F$ of a link $L$ in $S^3$, let $\alpha=(x_1,\ldots,x_\ell)$ be an ordered collection of non-intersecting, oriented, closed curves in $F$ that form an ordered basis for $H_1(F)$. Choose an orientation of $F$ by specifying a nowhere zero, normal vector field $n$ on $F$. For each curve $x_i$, let $x_i^+$ be the curve in $S^3$ that is $x_i$ pushed off the surface $F$ slightly in the direction of $n$. For each pair of curves $x_i,x_j^+$ we can compute their linking numbers (they do not intersect), and so we define the \textbf{Seifert matrix} of $F$ with orientation $n$ to be the matrix $[F] = (\lk(x_i,x_j^+))_{i,j=1}^\ell$. Of course this matrix depends on our choice of orientation of $F$ and basis for $H_1(F)$, so to include this dependence in the notation we may sometimes write $[F]_{\alpha,n}$. Suppose we have fixed our basis $x_1,\ldots,x_\ell$ with pushed off curves $x_1^+,\ldots,x_\ell^+$ relative to $n$. If we change the orientation of $F$ to $-n$, then we can translate $F$ a little bit so that $x_1^+,\ldots,x_\ell^+$ lie in $F$, in which case the curves $x_1,\ldots,x_\ell$ are now the pushed off curves relative to $-n$. So, the Seifert matrix with respect to this orientation is $(\lk(x_i^+,x_j))$. This is equal to $(\lk(x_j,x_i^+))$ by the symmetry of linking numbers, which is the transpose of the Seifert matrix computed with respect to $n$. So, reversing orientations transposed the Seifert matrix. If we fix the orientation, then any two bases of $H_1(F)$ are related by an invertible matrix $U$ (invertible over $\mathbb Z$ means $\det U=\pm 1$), so if $\alpha$ and $\beta$ are two bases related by $U$, then $[F]_\beta = U^T [F]_\alpha U$.
However, even if we consider two Seifert matrices $A$ and $B$ equivalent if they are related by $A=U^TBU$, for some invertible $U$, we do not get a knot invariant. We could just analyze how Seifert matrices transform under the Reidemeister moves, but a nicer way is to find a set of moves that can be performed on Seifert surfaces such that any two Seifert surfaces of a knot are related by these moves. On such move is \textbf{ambient isotopy}; that is, if $G : S^3 \times I \rightarrow S^3$ is an isotopy of $S^3$ and $F$ is a Seifert surface for a knot $k$, then we say $F' = G(F,1)$ is an ambiently isotopic Seifert surface of the isotopic knot $k' = G(k,1)$. Another move is to remove two disjoint discs $D_1,D_2$ in the interior of $F$, and glue in an embedded tube $S^1 \times I$ that is disjoint from the rest of $F$. This increases the genus of $F$ by one, but does not change the boundary. This move is called \textbf{stabilization}. The inverse of this move, called \textbf{destabilization}, is the process of finding a curve $c$ on $F$ such that $c$ bounds a disc in $S^3 \backslash F$, cutting $F$ open along $c$, and then closing $F$ back up by gluing two discs in. This move decreases the genus of $F$ by one.
\begin{prop}
\label{stabilization of Seifert surfaces}
Any two Seifert surfaces $F,F'$ of isotopic knots $k,k'$ can be connected via a finite sequence of ambient isotopies, stabilizations and destabilizations.
\end{prop}
\begin{proof}[Sketch]
Let $G : S^3 \times I \rightarrow S^3$ be an ambient isotopy between $k$ and $k'$, i.e. $G(k,1) = k'$. The set $M'$ of points $(G(x,t),t) \in S^3 \times I$, as $(x,t)$ ranges of $k \times I$, is an embedded surface in $S^3 \times I$. We can close this surface up by setting $M = (F \times 0) \cup M' \cup (F \times 1)$ to get an embedded, closed surface in $S^3$, and it bounds a 3-manifold $W$ ($\partial W = M$). Slightly perturbing $M$ and $W$, if necessary, we have that the time slices $S^3 \times t \cap W$ are smoothly embedded surfaces in $S^3$ with boundary $G(k,t)$, except possibly at finitely many points. From standard Morse theory that crossing these critical points changes $S^3 \times t \cap W$ by a surgery, depending on the index of the critical point. We can arrange that there are only critical points of index 1 and 2, which correspond to destabilization and stabilization respectively.
\end{proof}
We have already remarked that a change of basis of $H_1(F)$ corresponds to changing the Seifert matrix of $F$ by $U^TSU$, for some integral matrix with $\det U = \pm 1$. Let us see how stabilization of a Seifert surface changes the Seifert matrix. Let $F$ be a Seifert surface of $k$ with generators $x_1,\ldots,x_{2g}$ for $H_1(F)$. By forming $F' = F \sharp T'$ we are adding two generators to the first homology, one curve going around the longitude of $T'$ and one curve going around the meridian, so let $y$ and $z$ denote these curves. Since the torus $T'$ can knot and link with $F$ in complicated ways, we have that column and row of $[F']$ corresponding to $y$ can consist of arbitrary integers, except that either $\lk(y,z^+)=\pm 1$ and $\lk(z,y^+)=0$, or the opposite. Therefore the Seifert matrix $[F']$ is given by
\[ \begin{pmatrix} & & & * & 0 \\ & [F] & & \vdots & \vdots \\ & & & * & 0 \\ * & \cdots & * & * & \pm 1 \\ 0 & \cdots & 0 & 0& 0 \end{pmatrix} \text{\ \ \ \ \ or \ \ \ \ \ } \begin{pmatrix} & & & * & 0 \\ & [F] & & \vdots & \vdots \\ & & & * & 0 \\ * & \cdots & * & * & 0 \\ 0 & \cdots & 0 & \pm 1 & 0 \end{pmatrix} \]
This leads us to define two operations we can perform on integral matrices. Given an integral, $n \times n$ matrix $A$ (not necessarily invertible), performing an $S_1$ move is to change the matrix to $U^TAU$ for some invertible integral matrix $U$. Performing an $S_2$ move is to change the matrix to the $(n+2) \times (n+2)$ matrix given by
\begin{equation}
\label{S_2 move}
\begin{pmatrix} & & & * & 0 \\ & S & & \vdots & \vdots \\ & & & * & 0 \\ * & \cdots & * & * & \pm 1 \\ 0 & \cdots & 0 & 0& 0 \end{pmatrix}
\end{equation}
where the $*$'s can be any numbers. We say that two integral matrices are \textbf{$S$-related} one can be brought to the other with a finite sequence of $S_1$ and $S_2$ moves. An immediate consequence of \cref{stabilization of Seifert surfaces} is
\begin{cor}
\label{Seifert matrices are S-equivalent}
Any two Seifert matrices of a knot are $S$-equivalent.
\end{cor}
Therefore the $S$-equivalence class of any Seifert matrix of a knot is an invariant of the knot. There are simpler or more computable invariants we can derive from the Seifert matrix. Given a knot $k$, let $S$ be any Seifert matrix for $k$. The Laurent polynomial
\[ \Delta_k(t) = \det \left( t^{1/2} S - t^{-1/2} S^T \right) \]
in $t^{1/2}$ is called the \textbf{Alexander polynomial} of $k$. Technically we should call this the Conway normalization of Alexander's polynomial, as Alexander's original definition used the infinite cyclic covering of the knot complement, and was defined only up to multiplication by a unit. Since the dimension of $S$ is $2g$ (where $g$ is the genus of the surface associated to $S$) we have $\Delta_k(t)$ is actually a Laurent polynomial in $t$, and a direct substitution shows $\Delta_k(t)=\Delta_k(t^{-1})$. If $S$ is of genus $g$, then we can easily see that $\Delta_k(t) = t^{-g}\det(S-tS^T)$.
\begin{prop}
\label{Alexander polynomial is a knot invariant}
The Alexander polynomial does not depend on the Seifert matrix chosen, hence is a knot invariant.
\end{prop}
\begin{proof}
If $S'$ is another Seifert matrix for $k$, then $S$ and $S'$ are related by a finite sequence of $S_1$ and $S_2$ moves by \cref{Seifert matrices are S-equivalent}. Let $\Delta_k(t)$ denote the Alexander polynomial with respect to $S$, and $\Delta_k'(t)$ the Alexander polynomial with respect to $S'$. If $S' = U^TSU$, then we see that
\begin{align*}
\Delta_k'(t) &= \det \left( t^{1/2} U^TSU - t^{-1/2} U^TSU \right) \\
&= (\det U)^2 \det \left( t^{1/2} S - t^{-1/2} S^T \right) \\
&= \Delta_k(t)
\end{align*}
where the last line follows since $\det U = \pm 1$. On the other hand, if $S'$ is the result of performing the $S_2$ move on $S$ (see \eqref{S_2 move}), then $S'^T$ can be written as
\[ \begin{pmatrix} & & & * & 0 \\ & S^T & & \vdots & \vdots \\ & & & * & 0 \\ * & \cdots & * & * & 0 \\ 0 & \cdots & 0 & 1 & 0 \end{pmatrix} \]
and the matrix $t^{1/2} S' - t^{-1/2} S'^T$ can be written as
\[ \begin{pmatrix} & & & * & 0 \\ & t^{1/2} S - t^{-1/2} S^T & & \vdots & \vdots \\ & & & * & 0 \\ * & \cdots & * & * & t^{1/2} \\ 0 & \cdots & 0 & -t^{-1/2} & 0 \end{pmatrix} \]
When computing the determinant of this matrix we can perform a cofactor expansion along the last column, and then another cofactor expansion along the last row of the resulting matrix, in which case we see that the determinant of the above is precisely
\[ -t^{-1/2} \left( -t^{1/2} \det \left( t^{1/2} S - t^{-1/2} S^T \right) \right) = \Delta_k(t) \]
hence the Alexander polynomial is invariant under the $S_2$ moves.
\end{proof}
We can extend the definition of Alexander polynomial to links in the obvious way, but it will no longer be independent of the Seifert surface. The problem is that a link $L$ with $n$ components has $2^n$ orientations. A Seifert surface $F$ for $L$ can induce two of these orientations on $L$ (and they are opposite), and conversely we say that a Seifert surface is compatible with an oriented link if the induced orientation agrees with the link's orientation. While it is true that any two Seifert surfaces compatible with an oriented link are stably equivalent, it is \emph{not} true that any two Seifert surfaces of a link are stably equivalent. So, the Alexander polynomial is an invariant of oriented links with compatible Seifert matrices.
As mentioned above, the Alexander polynomial is not sensitive to the orientation of a knot, i.e. $\Delta_k(t) = \Delta_{-k}(t)$. Unfortunately, the Alexander polynomial is also not sensitive to changing a knot to its mirror. The \textbf{mirror} of a knot $k$ is the knot $\overline{k}$ obtained by reflecting $k$ through any plane in $\mathbb R^3$. This is equivalent to changing all over crossings to under crossings (and vice versa) in a diagram of $k$.
\begin{prop}
For any knot $k$, $\Delta_{k}(t) = \Delta_{\overline{k}}(t)$.
\end{prop}
\begin{proof}
If $S$ is a Seifert matrix of $k$, then $-S$ is a Seifert matrix of $\overline k$. Let $2m$ be the order of $S$. Then
\begin{align*}
\Delta_{\overline k}(t) &= \det\left( -t^{1/2} S + t^{-1/2} S^T \right) \\
&= \det\left( t^{-1/2} S^T - t^{1/2} S \right) \\
&= \det\left( t^{-1/2} S - t^{1/2} S^T \right) \\
&= \Delta_k(t^{-1}) = \Delta_k(t)
\end{align*}
\end{proof}
There are also scalar invariants of knots that one can derive from Seifert matrices.
\begin{prop}
\label{knot determinant}
If $S$ is the Seifert matrix of some Seifert surface of a knot $k$, then $|\det(S+S^T)|$ is a knot invariant.
\end{prop}
\begin{proof}
We just have to check that $|\det(S+S^T)|$ does not change after applying $S_1$ and $S_2$ moves to $S$. If we change $S$ by $U^TSU$, then clearly
\[ |\det(U^TSU + U^TS^TU)| = |(\det U)^2 \det(S+S^T)| = |\det(S+S^T)| \]
Let $S'$ be the matrix resulting from applying $S_1$ to $S$ (see \eqref{S_2 move}). Then $S'+S'^T$ is of the form
\[ \begin{pmatrix} & & & * & 0 \\ & S + S^T & & \vdots & \vdots \\ & & & * & 0 \\ * & \cdots & * & * & \pm 1 \\ 0 & \cdots & 0 & \pm 1 & 0 \end{pmatrix} \]
By adding multiples of the last column to the other columns, and multiples of the last row to the other rows we can bring this matrix into the form
\begin{equation}
\label{elementary transformations on S+S^T}
\begin{pmatrix} & & & 0 & 0 \\ & S + S^T & & \vdots & \vdots \\ & & & 0 & 0 \\ 0 & \cdots & 0 & 0 & \pm 1 \\ 0 & \cdots & 0 & \pm 1 & 0 \end{pmatrix}
\end{equation}
and the determinant of this matrix is $-\det(S+S^T)$. Therefore the absolute value of this determinant remains invariant under $S_1$ and $S_2$ moves.
\end{proof}
We call $|\det(S+S^T)|$ the \textbf{determinant} of the knot $k$, and denote it by $\det(k)$. The relationship between the determinant of $k$ and the Alexander polynomial is clear.
\begin{prop}
$|\Delta_k(-1)| = \det(k)$
\end{prop}
\begin{proof}
Let $S$ be any Seifert matrix of $k$ and let $2m$ be the order of $S$ (the order must be even for a knot). Then
\begin{align*}
|\Delta_k(-1)| &= \left| \det\left( (-1)^{1/2} S - (-1)^{-1/2} S^T \right) \right| \\
&= \left| \det\left( (-1)^{1/2} S + (-1)^{1/2} S^T \right) \right| \\
&= \left| (-1)^m \det(S+S^T) \right|
\end{align*}
\end{proof}
The Alexander polynomial satisfies a skein relation, which is closely related to another polynomial invariant that is essentially the same as the Alexander polynomial. For an oriented link $L$ and a diagram $D$ of $L$, choose any crossing in this diagram and let $L_+$ denote the diagram resulting from switching this crossing to a positive crossing, $L_-$ the diagram resulting from switching to a negative crossing, and $L_0$ the diagram resulting from resolving the crossing in the oriented way (see \cref{??}).
\begin{thm}
The Alexander polynomial of an oriented link $L$ satisfies the skein relation
\[ \Delta_{L_+}(t) - \Delta_{L_-}(t) = (t^{1/2}-t^{-1/2}) \Delta_{L_0}(t) \]
Further, this relation, along with the normalization $\Delta_\text{unknot}(t)=1$, uniquely determines the Alexander polynomial as a function from oriented links to $\mathbb Z[t^{1/2},t^{-1/2}]$.
\end{thm}
\begin{proof}
Let $F_0$ be the Seifert surface for $L_0$ constructed from Seifert's algorithm (see \cref{Seifert's algorithm}). We can construct Seifert surfaces $F_\pm$ for $L_\pm$ by gluing in a twisted band. Take a basis $\lcb x_1,\ldots,x_{k-1} \rcb$ for $H_1(F_0)$, then we can get a basis for $H_1(F_\pm)$ by adding one extra curve $x_k$ that goes through the twisted band once. Let $S_0$ be the Seifert matrix of $F_0$ computed with respect to $\lcb x_1,\ldots,x_{k-1} \rcb$. The Seifert matrices $S_\pm$ of $F_\pm$ are of the form
\[ S_+ = \begin{pmatrix}
& & & b_1 \\
& S_0 & & \vdots \\
& & & b_{k-1} \\
a_1 & \cdots & a_{k-1} & N
\end{pmatrix}
\ \ \ \ \ \ \
S_- = \begin{pmatrix}
& & & b_1 \\
& S_0 & & \vdots \\
& & & b_{k-1} \\
a_1 & \cdots & a_{k-1} & N-1
\end{pmatrix}
\]
for some integers $a_1,\ldots,a_k,b_1,\ldots,b_k,N$. Writing out the entries of $t^{1/2}S_\pm - t^{-1/2}S_\pm^T$ we see that the last column and row of these matrices are equal, except for the bottom-right corner entry where one matrix has $(t^{1/2}-t^{-1/2})N$ and the other has $(t^{1/2}-t^{-1/2})(N-1)$. So, if we compute $\Delta_{L_+}(t) - \Delta_{L_-}(t)$ by performing a cofactor expansion along the last columns, then the terms from $\Delta_{L_+}(t)$ will cancel those from $\Delta_{L_-}(t)$ except for the last term, which is equal to
\[ (-1)^{2k} (t^{1/2}-t^{-1/2})N \Delta_{L_0}(t) - (-1)^{2k} (t^{1/2}-t^{-1/2})(N-1) \Delta_{L_0}(t) = (t^{1/2}-t^{-1/2}) \Delta_{L_0}(t) \]
Therefore $\Delta_L(t)$ satisfies the skein relation.
\end{proof}
Although the Alexander polynomial satisfies a skein relation, this relation does not quite look like the relation for the Jones polynomial. To make things more uniform we introduce the Conway polynomial $\nabla_L(z)$. This is a function $\nabla : \lcb \text{oriented links} \rcb \rightarrow \mathbb Z[z]$ that is uniquely determined by the normalization $\nabla_{\text{unknot}}(z) = 1$ and the skein relation
\[ \nabla_{L_+}(z) - \nabla_{L_-}(z) = z \nabla_{L_0}(z) \]
The relationship between the Alexander and Conway polynomial is given by $\Delta_L(t) = \nabla_L\left( t^{1/2}-t^{-1/2} \right)$.
\subsection{Knot Signatures and 4-Ball Genus}
\label{Knot Signatures and 4-Ball Genus}
A very important invariant of knots is the signature.
\begin{prop}
\label{knot signature}
Let $\omega$ be a unit complex number not equal to 1, and let $S$ be a Seifert matrix of the knot $k$. Then the signature of the Hermitian matrix $(1-\omega)S + (1-\overline\omega)S^T$ over $\mathbb C$ is an invariant of the knot.
\end{prop}
\begin{proof}
We just have to check that the signature of $(1-\omega)S+(1-\overline\omega)S^T$ does not change after applying $S_1$ and $S_2$ moves to $S$. Clearly the signature is invariant under $S_1$ moves. Let $S'$ be the Seifert matrix resulting in applying an $S_2$ move, then we have $(1-\omega)S+(1-\overline\omega)S^T$ is congruent to
\[ \begin{pmatrix} & & & 0 & 0 \\ & (1-\omega)S + (1-\overline\omega)S^T & & \vdots & \vdots \\ & & & 0 & 0 \\ 0 & \cdots & 0 & 0 & 1-\omega \\ 0 & \cdots & 0 & 1-\overline\omega & 0 \end{pmatrix} \]
The signature of this block matrix is the sum of the signatures of the blocks. Since the signature of the bottom-right block is zero, we have that the signature of the above matrix is the signature of $(1-\omega)S+(1-\overline\omega)S^T$, hence the signature is invariant under $S_2$ moves.
\end{proof}
We call the signature of $(1-\omega)S+(1-\overline\omega)S^T$ the \textbf{$\omega$-signature} of the knot, and denote it by $\sigma_\omega(k)$. If $\omega=-1$, then we just call this \emph{the} signature of $k$, and denote it by $\sigma(k)$.
For a knot $k$ in $S^3$, consider the 4-ball $D^4$ bounded by $S^3$. Let $g^*(k)$ denote the minimal genus of a surface $F$ that can be smooth embedded in $D^4$ with $\partial F = k \subset S^3$. This knot invariant is called the \textbf{smooth 4-ball genus}, or sometimes the \textbf{smooth slice genus}. A knot is said to be \textbf{slice} if $g^*(k)=0$. The relationship between signature and slice knots is the following.
\begin{thm}
\label{slice knots and signature}
If $k$ is a slice knot, then $\sigma(k)=0$.
\end{thm}
We will outline the proof of this theorem in a series of lemmas. Suppose $k$ is a slice knot with slicing disc $D \subset D^4$. If $F$ is a Seifert surface for $k$ in $S^3$, then we can glue together $F \cup D$ to obtain a closed surface in $D^4$.
\begin{lem}
The surface $F \cup D$ bounds a smoothly embedded, orientable 3-manifold $M$ in $D^4$ with $M \cap S^3 = F$.
\end{lem}
\begin{lem}
Let $M$ be a compact, orientable 3-manifold such that $\partial M$ is a surface of genus $g$. Then the kernel of the map $i_* : H_1(\partial M;\mathbb Q)\rightarrow H_1(M;\mathbb Q)$ induced by the inclusion is of dimension $g$.
\end{lem}
\begin{proof}
We can put the homology and cohomology long exact sequences of the pair $(M,\partial M)$ into the following commutative diagram
\[
\xymatrix
{
H_2(M,\partial M;\mathbb Q) \ar[r] \ar[d] & H_1(\partial M;\mathbb Q) \ar[r]^{i_*} \ar[d] & H_1(M;\mathbb Q) \ar[d] \\
H^1(M;\mathbb Q) \ar[r]^{i^*} & H^1(\partial M;\mathbb Q) \ar[r]^{\delta} & H^2(M,\partial M;\mathbb Q)
}
\]
where the vertical arrows are the Poincar\'{e} and Poincar\'{e}-Lefschetz duality isomorphisms. By the universal coefficient theorem we have that $H^1(\partial M;\mathbb Q)$ is dual to $H_1(\partial M;\mathbb Q)$, $H_1(M;\mathbb Q)$ is dual to $H^1(M;\mathbb Q)$, and $i^*$ is dual to $i_*$. By the rank-nullity theorem we have
\[ 2g = \dim \image i_* + \dim \ker i_* = \dim \image i_* + \dim \ker \delta = \dim \image i_* + \dim \image i^* = 2 \dim \image i_* \]
hence $\dim \image i_* = g$, and so $\dim \ker i_* = g$.
\end{proof}
\begin{lem}
There is a basis $\lcb x_i \rcb$ over $\mathbb Z$ for $H_1(\partial M;\mathbb Z)$ such that $i_*(x_j)=0$ in $H_1(M;\mathbb Q)$ for $1 \leq j \leq g$.
\end{lem}
\begin{proof}
We think of $H_1(\partial M) = \mathbb Z^{2g}$ as sitting inside $H_1(\partial M;\mathbb Q) = \mathbb Q^{2g}$. Let $U$ be the $g$ dimensional subspace of $\mathbb Q^{2g}$ which is the kernel of the map $\mathbb Q^{2g} \rightarrow H_1(M;\mathbb Q)$ induced by the inclusion. We can take a basis $\lcb x_1,\ldots,x_g \rcb$ for this vector space such that $x_i \in \mathbb Z^{2g}$, and so let $\widetilde{U}$ be the $\mathbb Z$-span of these vectors.
As a $\mathbb Z$-module we can write $\mathbb Z^{2g} / \widetilde U$ as $A / \widetilde U \oplus B / \widetilde U$, where $A$ and $B$ are submodules of $\mathbb Z^{2g}$ such that $A / \widetilde U$ is free and $B / \widetilde U$ is torsion. The fact that this latter module is torsion means that for $b \in B$ there is an integer $n$ such that $nb \in \widetilde U$, hence $nb \in U$, or simply $b \in U$. Thus a $\mathbb Z$-basis for $B$ also serves as a $\mathbb Q$-basis for $U$. So, the $\mathbb Z$-basis we take for $H_1(\partial M)$ is any extension of a basis for $B$ to a full basis for $H_1(\partial M)$.
\end{proof}
\begin{lem}
Let $F$ be a genus $g$ Seifert surface for a slice knot $k$. Then there is a basis for $H_1(F)$ such that the associated Seifert matrix is of the form
\[ S = \begin{pmatrix} 0 & P \\ Q & R \end{pmatrix} \]
for some $g \times g$ matrices $P,Q,R$.
\end{lem}
\begin{proof}
Let $D$ be a slicing disc for $k$ in $D^4$ so that there is a 3-manifold $M$ with $\partial M = F \cup D$. Let $\lcb x_i \rcb$ be a $\mathbb Z$-basis for $H_1(\partial M)$ as in the previous lemma so that $x_i=0$ in $H_1(M;\mathbb Q)$ for $1 \leq i \leq g$. We can assume that the homology classes $x_i$ are represented by oriented, closed curves in $\partial M$, and we will also write $x_i$ to denote these curves. For each $1 \leq i \leq g$ we can find a non-zero integer $n_i$ such that $n_i x_i = 0$ in $H_1(M)$. The class $n_i x_i$ can be represented by a closed curve in $\partial M$, which we also denote by $n_i x_i$, and it bounds a 2-chain in $M$. This means we can find maps $f_i : F_i \rightarrow D^4$, where $F_i$ is a surface with boundary, such that $f_i(\partial F_i) = n_i x_i$. If we push the curve $n_i x_i$ off of $\partial M$ in the normal direction, then we can also pushing the mapping $f_i$ in this direction. Therefore, for $1 \leq i,j, \leq g$ the curves $n_i f_i$ and $(n_j f_j)^+$ bound disjoint surfaces in $D^4$, and hence
\[ 0 = \lk(n_if_i,(n_jf_j)^+) = n_in_j \lk(f_i,f_j^+) \]
Therefore the upper-left $g \times g$ block of the Seifert matrix of $F$ with respect to the basis $\lcb x_i \rcb$ is zero.
\end{proof}
\begin{lem}
If $B$ is a symmetric, non-degenerate bilinear form on an even dimensional space such that there is a half dimensional space $U$ with $B(u,u')=0$ for all $u,u' \in U$, then $B$ has zero signature.
\end{lem}
\begin{proof}
Suppose the space $V$ that $B$ is defined on is of dimension $2g$, and let $U$ be a subspace of dimension $g$ where $B$ vanishes. If $V^\pm$ denote the maximal positive/negative-definite subspaces of $V$ with respect to $B$, then $V = V^+ \oplus V^-$ since $B$ is non-degenerate. However, we also have $U \cap V^+ = \lcb 0 \rcb$ and $U \cap V^- = \lcb 0 \rcb$, hence $\dim V^+ \leq g$ and $\dim V^- \leq g$. Since we need $\dim V^+ + \dim V^- = 2g$ we are forced to have $g = \dim V^+ = \dim V^-$, which implies $\sigma(B)=0$.
\end{proof}
\begin{proof}[Proof of \Cref{slice knots and signature}]
Clearly the combination of the previous lemmas proves the theorem.
\end{proof}
More generally one can also prove
\begin{thm}
\label{signature 4-ball genus bound}
For any knot $k$ we have $|\sigma(k)| \leq 2g^*(k)$.
\end{thm}
For a knot $k$, let $u(k)$ denote the minimum number of crossing changes needed in a diagram of $k$ in order to obtain the unknot. This number is called the \textbf{unknotting number} of $k$. We can think of the process of unknotting $k$ as a movie in $S^3 \times [0,1]$ (i.e. an embedded surface), such that the $t$ slice of the surface is an embedded knot in $S^3 \times \lcb t \rcb$ for all but finitely many values of $t$. At one of these singular times $t$ we can find $\epsilon > 0$ small enough so that the $t \pm \epsilon$ slices of the surface are embedded knots, and differ only by a crossing change. Further, if there are two times $t_0,t_1$ such that the slices of the surface at all times between $t_0,t_1$ are embeddings, then the slice at $t_0$ and $t_1$ differ only by Reidemeister moves.
Now, let $\Sigma \subset S^3 \times [0,1]$ be an embedded surface such that $\Sigma \cap S^3 \times \lcb 0 \rcb = k$ and $\Sigma \cap S^3 \times \lcb 1 \rcb$ is the unknot, and let $\Sigma_t = \Sigma \cap S^3 \times \lcb t \rcb$ and $\Sigma_{\leq t} = \Sigma \cap S^3 \times [0,t]$. If $\Sigma_t$ is singular, then we claim that the genus of $\Sigma_{\leq t+\epsilon}$ is strictly greater than $\Sigma_{\leq t-\epsilon}$. \unfinished .... We have proved the following.
\begin{lem}
$g^*(k) \leq u(k)$
\end{lem}
\begin{cor}
$|\sigma(k)| \leq 2g^*(k) \leq 2u(k)$
\end{cor}
We say that two knots $k_1,k_2$ in $S^3$ are \textbf{concordant} if there is a smooth embedding of a cylinder $S^1 \times I$ in $S^3 \times I$ such that $S^1 \times \lcb i \rcb$ maps to $k_i \times \lcb i \rcb$ ($i=0,1$). Concordance defines an equivalence relation on the set of knots in $S^3$. If a knot $k$ is concordant to the unknot, then we can cap off the unknot with a disc and obtain a smoothly embedded disc in $D^4$ that is bounded by $k$, hence $k$ is slice. Therefore a knot is concordant to the unknot if and only if it is slice.
Let $\mathcal C_1$ denote the concordance classes of knots in $S^3$. If $k_1,k_1'$ and $k_2,k_2'$ are concordant pairs of knots, then $k_1 \# k_2$ and $k_1' \# k_2$ are concordant. The cobordism can be constructed by splicing together the cylinders connected $k_1$ to $k_1'$ and $k_2$ to $k_2'$. Next we claim that for any knot $k$, the knot $k \# \overline k$ is slice. Define the following subsets of $\mathbb R^4$:
\[ \mathbb R^3 = \lcb (x,y,z,0) \in \mathbb R^4 \rcb \]
\[ \mathbb R_-^3 = \lcb (x,y,z,0) \in \mathbb R^4 \st z \leq 0 \rcb \]
\[ \mathbb R_+^3 = \lcb (x,y,z,0) \in \mathbb R^4 \st z \geq 0 \rcb \]
\[ \mathbb R^2 = \lcb (x,y,0,0) \in \mathbb R^4 \rcb \]
Arrange the knot $k \# \overline k$ in $\mathbb R^3$ so that its intersection with the plane $\mathbb R^2$ in $\mathbb R^4$ consists of two points on the band used in the connect sum. Let $k^+$ denote the intersection of $k \# \overline k$ with $\mathbb R_+^3$, which is just an arc, so that if $k_-$ is the reflection of $k_+$ through $\mathbb R^2$, then $k_+ \cup k_-$ is isotopic to $k \# \overline k$. Consider the set of points
\[ D = \lcb (x,y,z\cos\theta,z\sin\theta) \in \mathbb R^4 \st (x,y,z,0) \in k_+, 0 \leq \theta \leq \pi \rcb \]
That is, we spin all the points on $k_+$ about the plane $\mathbb R^2$ in $\mathbb R^4$ through all angles from $0$ to $\pi$. Clearly $D$ is just a disc (being a half spin of an arc) with boundary $k \# \overline k$, and this provides us with a slicing disc for $k \# \overline k$.
We can now conclude that $\mathcal C_1$ forms a group under connect sum with $[\text{unknot}]$ serving as the identity and the mirror of a knot serving as its inverse. Further, if $k_1$ and $k_2$ are concordant, then $0=\sigma(k_1 \# \overline k_2) = \sigma(k_1)-\sigma(k_2)$, hence $\sigma$ descends to a well-defined group homomorphism $\sigma : \mathcal C_1 \rightarrow \mathbb Z$.
A knot $k$ is said to be amphichiral if it is isotopic to its mirror $\overline k$. In particular, such a knot is an element of order 2 in $\mathcal C_1$. The figure-eight know is amphichiral, as shown in \cref{figure-eight-amphichiral}.
\begin{figure}[tb]
\centering
\includegraphics[scale=1]{graphics/figure-eight-amphichiral}
\caption{Showing the figure-eight knot is isotopic to its mirror}
\label{figure-eight-amphichiral}
\end{figure}
\subsection{Cyclic and Branched Covers}
\label{Cyclic and Branched Covers}
Now let us discuss Alexander's original definition of his polynomial using the infinite cyclic cover of the complement of a link. To construct this covering space we fix an oriented link $L$, set $X = S^3 \backslash L$, and fix a Seifert surface $F$ (compatible with the orientation of $L$). Let $Y$ be the space obtained from cutting $X$ along $F$ so that $\partial Y$ consists of two disjoint copies of $F$, denoted by $F_+$ and $F_-$; that is, $Y$ is homeomorphic to the complement of an open tubular neighborhood of $F$ in $S^3$. There is a natural identification $\phi : F_- \rightarrow F_+$. Take the countable collection $Y \times \lcb i \rcb$ ($i \in \mathbb Z$) of copies of $Y$, and let $h_i : Y \rightarrow Y_i$ be the obvious homeomorphism. Form the space $X_\infty$ by gluing $Y_i$ to $Y_{i+1}$ via the homeomorphism $h_{i+1} \circ \phi \circ h_i^{-1}$; that is, we glue $Y_i$ to $Y_{i+1}$ by gluing $F_- \times \lcb i \rcb$ to $F_+ \times \lcb i+1 \rcb$ via the map $h_{i+1} \circ \phi \circ h_i^{-1}$. The obvious projection $\pi : X_\infty \rightarrow X$, where $\pi|_{Y_i} : Y_i \rightarrow X$ is given by $\pi(x,i)=x$, is an infinite cyclic covering space of $X$.
Let $t$ denote the covering transformation of $X_\infty$ whose restriction $t|_{Y_i} : Y_i \rightarrow X_\infty$ is given by $t(x,i) = (x,i+1)$. This map induces an isomorphism (which we also denote by $t$) $t : H_1(X_\infty) \rightarrow H_1(X_\infty)$, turning $H_1(X_\infty)$ into a $\mathbb Z[t,t^{-1}]$-module, called the Alexander module of $L$. Of course, we need to show that this module is independent of the various choices in its construction
\begin{prop}
If $X_\infty$ and $X_\infty'$ are the infinite cyclic coverings of the link complement $X$ of a link $L$ constructed from two different Seifert surfaces, then $X_\infty$ and $X_\infty'$ are isomorphic has covering spaces, and $H_1(X_\infty)$ and $H_1(X_\infty')$ are isomorphic as $\mathbb Z[t,t^{-1}]$ modules.
\end{prop}
\begin{proof}
Recall that covering spaces are determined up to equivalence by the image of the fundamental group of the total space under the covering projection. That subgroup of the fundamental group of the base space consists of those loops in the base that lift to loops in the total space. So, we show that the notion of lifting loops to loops does not depend on the Seifert surface $F$ used in the construction of $X_\infty$.
Let $\alpha : [0,1] \rightarrow X$ be any path in $X$ and consider a lift $\tilde\alpha : [0,1] \rightarrow X_\infty$. This path is a loop in $X_\infty$ if and only if $\tilde\alpha(0)$ and $\tilde\alpha(1)$ are both in $X_\infty$, which means the path intersects the chambers $F_\pm \times \lcb i \rcb$ with algebraic multiplicity zero. Down in $X$ this means that that $\alpha$ has zero algebraic intersection number with $F$, which we know is equivalent to $\alpha$ and $L$ having zero linking number (which is the sum of the linking numbers with each component of $L$). This statement is of course independent of $F$, hence our notion of path lifting is independent of $F$, and so these covering spaces are equivalent.
Next we have to check that any two coverings are equivalent via a $t$-equivariant homeomorphism in order to ensure that the first homology $\mathbb Z[t,t^{-1}]$-modules are isomorphic. Suppose $p' : X_\infty' \rightarrow X$ is another covering space constructed with a different Seifert surface $F'$. Then there is a homeomorphism $h : X_\infty \rightarrow X_\infty'$ such that $p' \circ h = p$. Let $\alpha$ be a path in $X_\infty$ from a point $a$ to the translate $t \cdot a$. In the base, the path $p \circ \alpha$ is a loop with linking number 1 with $L$. Then $p' \circ h \circ \alpha = p \circ \alpha$, so $h \circ \alpha$ is a path in $X_\infty'$ from a point $a'$ to its translate $t \cdot a'$, hence $t \cdot h(a) = h(t \cdot a)$ and so $h$ is $t$-equivariant.
\end{proof}
Since the covering space $X_\infty \rightarrow X$ is well-defined given a link $L$ we have that the $\mathbb Z[t,t^{-1}]$-module $H_1(X_\infty)$ is an invariant of the link, called the \textbf{Alexander module}. Let us connect this invariant with the Alexander polynomial defined earlier. Given an $R$-module $M$, a presentation of $M$ is an exact sequence
\[ F \stackrel{\alpha}{\longrightarrow} E \stackrel{\phi}{\longrightarrow} M \rightarrow 0 \]
where $F$ and $E$ are free $R$-modules. If we fix bases for $F$ and $E$ then we get a matrix representation of $\alpha$, called the presentation matrix of $M$. The $r$-th elementary ideal of $M$ is the ideal of $R$ generated by all $(m-r+1) \times (m-r+1)$ minors of a presentation matrix of $M$. In particular, the first elementary ideal of $M$ is simply the determinant of any presentation matrix of $M$. In order to state the relationship between the Alexander polynomial and the Alexander module, we first describe another way of viewing Seifert matrices.
\begin{lem}
Let $F$ be a compact, connected, orientable surface with boundary smoothly embedded in $S^3$. Then $H_1(S^3 \backslash F)$ is isomorphic to $H_1(F)$, and there is a unique non-singular bilinear form
\[ \beta : H_1(F) \times H_1(S^3 \backslash F) \rightarrow \mathbb Z \]
with the property that $\beta([c],[d])=\lk(c,d)$ for any oriented, simple closed curves $c,d$ in $F$ and $S^3\backslash F$ respectively.
\end{lem}
\begin{proof}
Suppose $F$ is of genus $g$ with $n$ boundary componponents. Let $U$ be a tubular neighborhood of $F$ in $S^3$ (which is just a handlebody of genus $2g+n-1$), and $V$ the closure of its complement so that $U \cap V = \partial U$ (which is a surface of genus $2g+n-1$). From the Mayer-Vietoris sequence we have
\[ 0 = H_2(S^3) \longrightarrow H_1(U \cap V) \longrightarrow H_1(U) \oplus H_1(V) \longrightarrow H_1(S^3) = 0 \]
Therefore the middle map is an isomorphism and since $H_1(U \cap V) \cong \mathbb Z^{2(2g+n-1)}$ and $H_1(U) \cong \mathbb Z^{2g+n-1}$, we must have $H_1(S^3 \backslash F) = H_1(V) \cong \mathbb Z^{2g+n-1}$. Pick a basis $\lcb x_i,y_i \rcb_{i=1}^{2g+n-1}$ of oriented, simple, closed curves for $H_1(U \cap V)$ such that $\lcb x_i \rcb$ and $\lcb y_i \rcb$ are bases for $H_1(F)$ and $H_1(S^3 \backslash F)$ respectively, with the property that $y_i$ bounds a disc in $U$ and $x_i$ intersects only the disc of $y_i$ in precisely one point. Also choose these curves such that $\lk(x_i,y_i)=\delta_{ij}$.
Now define the map $\beta : H_1(F) \times H_1(S^3 \backslash F) \rightarrow \mathbb Z$ on the bases by $\beta(x_i,y_j) = \delta_{ij}$, and extend linearly. Let $c$ and $d$ be simple, closed curves in $F$ and $S^3 \backslash F$ such that $[c] = \sum a_i x_i$ and $[d] = \sum b_j y_j$. Then $\lk(x_i,d)$ is an integer $n$ such that $[d] = n \mu_i \in H_1(S^3 \backslash x_i)$, where $\mu_i$ is a meridian of $x_i$. It is clear from construction that $y_i$ is a meridian of $x_i$, hence
\[ \sum b_j y_j = [d] = n \mu_i = n y_i \]
hence $n = b_i$. On the other hand, $\lk(c,d)$ is an integer $m$ such that $[c] = m \mu \in H_1(S^3 \backslash d)$, where $\mu$ is a meridian of $d$. Since $[c] = \sum a_i x_i$ and $\lk(x_i,d) = b_i$, we have that $x_i$ contributes $b_i$ to the coefficient of $\mu$ in $H_1(S^3 \backslash d)$. Therefore $\lk(c,d) = \sum a_i b_i$, which is precisely $\beta([c],[d])$.
\end{proof}
Now give $F$ an orientation via a normal, non-vanishing vector field $n$, and let $i^+,i^- : F \rightarrow S^3 \backslash F$ be the maps that slightly push the points in the direction of $n$ and $-n$, respectively. For a homology class $x \in H_1(F)$, let $x^\pm$ be the class $i_*^\pm x$ in $H_1(S^3 \backslash F)$. Define the pairing $\alpha : H_1(F) \times H_1(F) \rightarrow \mathbb Z$ by $\alpha(x,y) = \beta(x,y^+)$. This is called the \textbf{Seifert form} of $F$. For a basis $\lcb x_i \rcb$ of $H_1(F)$ we can form the matrix $S = (\alpha(x_i,x_j))_{ij}$, called the Seifert matrix of $F$.
The relationship between the Alexander polynomial and Alexander module can be stated as the following.
\begin{thm}
If $S$ is a Seifert matrix for a knot $k$ in $S^3$, then $tS-S^T$ is a presentation matrix for $H_1(X_\infty)$ as a $\mathbb Z[t,t^{-1}]$-module.
\end{thm}
\begin{proof}
In the notation we used when constructing $X_\infty$, let us write $X_\infty = U \cup V$, where $U = \cup_i Y_{2i}$ and $V = \cup_i Y_{2i+1}$. We have $H_0(U \cap V) = \oplus_i H_0(F_i)$, and the action of $t$ on this group shifts the summands of $\mathbb Z$ by one, so we can identify $H_0(U \cap V)$ with $\mathbb Z[t,t^{-1}] \otimes H_0(F)$ generated by $1 \otimes 1$. Similarly, $H_0(U) \oplus H_0(V) = \oplus_i H_0(Y_i)$ and so we can identify this with $\mathbb Z[t,t^{-1}] \otimes H_0(Y)$ generated by $1 \otimes 1$. With these identifications we see that the map $H_0(U \cap V) \rightarrow H_0(U) \oplus H_0(V)$ in the Mayer-Vietoris sequence is given by $1 \otimes 1 \mapsto -1\otimes 1 + t \otimes 1$, and so is injective. Therefore the map $H_1(U) \oplus H_1(V) \rightarrow H_1(X_\infty)$ is the Mayer-Vietoris sequence is surjective.
Let is now fix a basis $\lcb x_i \rcb$ for $H_1(F)$ and a dual basis $\lcb y_i \rcb$ with respect to the bilinear form $\beta$. Let $S = (s_{ij})$ be the Seifert matrix of $F$ with respect to this basis. Similar to the above arguments we have an identification of $H_1(U \cap V)$ with $\mathbb Z[t,t^{-1}] \otimes H_1(F)$, the latter being generated by $\lcb 1 \otimes x_i \rcb$, and an identification of $H_1(U) \oplus H_1(V)$ with $\mathbb Z[t,t^{-1}] \otimes H_1(Y)$, the latter being generated by $\lcb 1 \otimes y_i \rcb$. With these identifications, the map $H_1(U \cap V) \rightarrow H_1(U) \oplus H_1(V)$ in the Mayer-Vietoris sequence is given by $1 \otimes x_i = -1 \otimes x_i^- + t \otimes x_i^+$. However, as elements in $H_1(S^3 \backslash F)$, the cycles $x_i^+$ and $x_i^-$ can be written as linear combinations of the $y_j$'s. To figure out the coefficients let $x_i^+ = \sum a_j y_j$, then applying $\beta(x_k,-)$ to both sides yields
\[ \beta(x_k,x_i^+) = \beta(x_k,\sum a_j y_j) = a_k \]
where the left side is simply $s_{ki}$. Therefore
\[ x_i^+ = \sum_j s_{ji} y_j \]
\[ x_i^- = \sum_j s_{ij} y_j \]
We now have
\[ -1 \otimes x_i^- + t \otimes x_i^+ = \sum_j (-1 \otimes s_{ij} y_j + t \otimes s_{ji} y_j) \]
With, with respect to our bases we have that the matrix of the map in the Mayer-Vietoris sequence is given by $tS-S^T$.
\end{proof}
We can change the construction of the infinite cyclic cover of the link exterior slightly to obtain a finite cyclic cover. For an integer $k > 0$, take $k$ copies of $Y$ (the link exterior cut open along a Seifert surface) by setting $Y_i = Y \times \lcb i \rcb$ ($i=0,1,\ldots,k-1$). For the space $X_k$ by gluing $Y_i$ to $Y_{i+1}$ via the homeomorphism $h_{i+1} \circ \phi \circ h_i^{-1}$ (where all induces are taken modulo $k$). Then the obvious projection $\pi_k : X_k \rightarrow X$ is a $k$-fold cyclic covering of the link exterior. Note that we have an infinite cyclic covering $X_\infty \rightarrow X_k$ that maps the points in $Y_i \subset X_\infty$ identically onto $Y_{i \modulo k}$. In fact, the space $X_k$ is simply the quotient $X_\infty / \< t^k \>$.
This collection of covering spaces depends only on the link $L$ and not on the various choices used to define it.
This collection of covering spaces can now be extending to cyclic branched covers $\tilde X_k \rightarrow S^3$ of the 3-sphere, branched over the link $L$. A loop in $X$ lifts to a loop in $X_k$ if and only if it has linking number 0 (modulo $k$) with $L$. For a component $L_i$ from the link $L$, let $\nu L_i$ be a tubular neighborhood. We can identify $\partial \nu L_i$ with $S^1 \times S^1$ by choosing a meridian $\mu_i$ and longitude $\ell_i$ of $\partial \nu L_i$ such that $\ell_i$ links zero times with $L_i$. Then $\mu_i^k$ (the $k$-th power of $\mu_i$) and $\ell_i$ lift to a loops in $X_k$ which identify a torus in $X_k$. The restriction of the covering map to this torus gives a $k$-fold covering of one torus over another which is equivalent to $(z,w) \mapsto (z,w^k)$. We can extend this to a branched covering $S^1 \times D^2 \rightarrow S^1 \times D^2$ given by $(z,w)=(z,w^k)$, where this is branched over the core circle $S^1 \times 0$. So, if we glue solid tori into the boundary components of $X_k$ we obtain a space $\tilde X_k$ with a natural projection $\tilde X_k \rightarrow S^3$ that is a $k$-fold covering branched over $L$.
\subsection{Fox Calculus and the Alexander Polynomial}
\label{Fox Calculus and the Alexander Polynomial}
Now we describe a method to compute the Alexander polynomial from the knot group via Fox's free differential calculus. Consider a finitely presented group $G = \< x_1,\ldots,x_n \st r_1,\ldots,r_m \>$, and let $W$ be the universal cover of a bouquet of $n$ circles, which correspond to the generators $x_i$. Let $X_i$ be the edge of $W$ corresponding to the lift of the loop $x_i$. Then every edge in $W$ can be identified as a translate of the "basis" $\lcb X_i \rcb$, hence any word $\omega$ in the $x_i$'s can be lifted to a formal sum of simplices in $W$ with coefficients in $G$, which we will denote by $\widetilde\omega$. For example, the lift of the word $x_1x_2$ is the formal sum $X_1 + x_1 X_2$. More generally, a word $\omega$ written as the product of two other words $\omega = \omega_1\omega_2$ can be lifted as $\widetilde{\omega_1\omega_2} = \widetilde{\omega_1} + \omega_1 \widetilde{\omega_2}$.
For example, consider the word $\omega = x^2 y x^{-1}$ in a group with two generators $x,y$. Then
\begin{align*}
\widetilde{x^2 y x^{-1}} &= \widetilde{x} + x \widetilde{xyx^{-1}} \\
&= \widetilde{x} + x \widetilde{x} + x^2 \widetilde{yx^{-1}} \\
&= \widetilde{x} + x \widetilde{x} + x^2 \widetilde{y} + x^2 y \widetilde{x^{-1}}
\end{align*}
This last line is precisely the formal simplex $(1+x+x^2yx^{-1})X + x^2 Y$. We denote the coefficient of $X$ in $\widetilde{\omega}$ by $\pfrac{\omega}{x}$, and the coefficient of $Y$ by $\pfrac{\omega}{y}$. Note that these coefficients of are elements of the group ring $\mathbb Z[G]$.
More generally, for a word $\omega$ in the $x_i$'s of a group $G = \< x_1,\ldots,x_n \st r_1,\ldots,r_m \>$ we let $\pfrac{\omega}{x_i}$ be the coefficient of $X_i$ in $\widetilde{\omega}$, i.e.
\[ \widetilde{\omega} = \pfrac{\omega}{x_1} X_1 + \cdots + \pfrac{\omega}{x_n} X_n \]
These formalities have the following properties
\begin{enumerate}
\item $\displaystyle \pfrac{(\omega \tau)}{x} = \pfrac{\omega}{x} + \omega \pfrac{\tau}{x}$
\item $\displaystyle \pfrac{(1)}{x} = 0$ and $\displaystyle \pfrac{(\omega^{-1}}{x} = -\omega^{-1} \pfrac{\omega}{x}$
\item $\displaystyle \pfrac{x^n}{x} = \begin{cases} 1+x+x^2+\cdots+x^{n-1} & n>0 \\ -x^{-1}-x^{-2}-\cdots-x^{n} & n < 0 \end{cases}$
\end{enumerate}
We can now state Fox's algorithm for computing the Alexander polynomial. Present the knot group of $k$ as $\pi_1(S^3 \backslash k) = \< x_1,\ldots,x_n \st r_1,\ldots,r_m \>$, and form the Jacobian matrix $J = \left( \pfrac{r_i}{x_j} \right)$. Let $\phi : \pi_1(S^3 \backslash k) \rightarrow H_1(S^3 \backslash k) = \< t \>$ be the abelianization map (we can think of $t$ as the meridian of $k$), and let $J^\phi$ be the image of the Jacobian under this map. Then $\Delta_k(t)$ is any generator of the ideal generated by \emph{all} maximal minors of $J^\phi$ (up to sign and multiples of $t$).
Let us apply this algorithm to compute the Alexander polynomial of the torus knot $T_{p,q}$. As we have seen the knot group is isomorphic to $\< x,y \st x^p = y^q \>$. First we figure out the abelianization map $\phi$. Since we can think of $t$ as the meridian of $T_{p,q}$ it is clear that $\phi(x)=t^q$ and $\phi(y)=t^p$. Let $r=x^py^{-q}$, then we have
\begin{align*}
\pfrac{r}{x} &= \pfrac{x^p}{x} + x^p \pfrac{y^{-q}}{x} = 1+x+x^2 + \cdots + x^{p-1} \\
\pfrac{r}{y} &= \pfrac{x^p}{y} + x^p \pfrac{y^{-q}}{y} = -x^p \left( y^{-1} + y^{-2} + \cdots + y^{-q} \right)
\end{align*}
and their abelianizations
\begin{align*}
\pfrac{r}{x}^\phi &= 1+t^q+x^{2q} + \cdots + t^{q(p-1)} = \frac{1-t^{pq}}{1-t^q} \\
\pfrac{r}{y}^\phi &= -t^{pq} \left( t^{-p} + t^{-2p} + \cdots + t^{-pq} \right) = \frac{-t^{pq}t^{-p}(1-t^{-pq})}{1-t^{-p}}
\end{align*}
These polynomials generate the same ideal generated by the following two polynomials
\[ \frac{1-t^{pq}}{1-t^q} \ \ \ \ \ \ \ \ \frac{1-t^{pq}}{1-t^p} \]
We want to find the $\gcd$ of these polynomials. Since $p$ and $q$ are relatively prime we have that $1-t$ is the $\gcd$ of $1-t^q$ and $1-t^p$, and so using the fact that $\gcd(a/b,a/c) = a \gcd(b,c) / (bc)$ we now have that a generator for the ideal generated by the polynomials given above is
\[ \frac{(1-t)(1-t^{pq})}{(1-t^p)(1-t^q)} \]
\subsection{Applications}
\label{Knot Applications}
Seifert matrices are closely related to the intersection form on a compact, orientable surface. Intersection forms are defined on any even dimensional manifold, and we give a fuller treatment for 4-manifolds in \cref{Intersection Forms on 4-Manifolds}, so we will just briefly state the facts. For a closed, oriented surface $M$, Poincar\'{e} duality gives us a non-degenerate, antisymmetric, bilinear form $Q_M : H^1(M) \otimes H^1(M) \rightarrow \mathbb Z$ defined by $Q_M(\alpha,\beta) = (\alpha \smallsmile \beta,[M])$. It follows that the matrix representation of $Q_M$ in any basis of $H^1(M)$ has determinant $\pm 1$. However, we have the following lemma concerning these types of matrices.
\begin{lem}
If $A$ is the matrix representation of a non-degenerate, antisymmetric bilinear form, then there is a matrix $U$ such that $A = U^T J U$, where $J$ is a block diagonal matrix of copies of
\[ \left( \begin{array}{rr} 0&1\\-1&0 \end{array} \right) \]
\end{lem}
Therefore the matrix of $Q_M$ with respect to any basis must have determinant $+1$. There is a more geometric definition of the intersection form on $M$. A homology class $a \in H_1(M)$ can be represented by a closed, smoothly embedded 1-dimensional submanifold of $M$, call it $C_a$. If $a,b \in H_1(M)$ with representative submanifolds $C_a$ and $C_b$, then perturb them a little so that they intersect transversely, and let $a \cdot b$ be the signed sum of their points of intersection, called the intersection product. Then one can show that $Q_M(\alpha,\beta)=PD(\alpha) \cdot PD(\beta)$.
If $M$ has one boundary component (i.e. $M$ is a Seifert surface of some knot), then $Q_M$ is still a non-degenerate, antisymmetric, bilinear form, and it can still be expressed in terms of intersections of homology classes. If $M$ has more than one boundary component, then $Q_M$ is an antisymmetric, bilinear form, and can be defined in terms of intersections, but it is now necessarily degenerate, and so any matrix representation has zero determinant.
\begin{prop}
\label{intersection forms and Seifert matrices}
Let $F$ be a Seifert surface of a knot $k$ with Seifert matrix $S$ in some basis of $H_1(F)$. Then the intersection form of $F$ in this basis is precisely $S-S^T$.
\end{prop}
\begin{proof}
At a positive crossing of $x_i$ and $x_j$ we see that $x_j^+$ is above $x_i$. So we have $\lk(x_i,x_j^+)$ contributes $+1$ at this crossing and $\lk(x_j,x_i^+)$ does not contribute anything (since we only count crossings where the first strand passes under). On the other hand, at a negative crossing of $x_i$ and $x_j$ we have that $x_j^+$ passes below $x_i$. Hence $\lk(x_i,x_j^+)$ contributes nothing at this crossing and $\lk(x_j,x_i^+)$ contributes $+1$. Therefore $x_i \cdot x_j = \lk(x_i,x_j^+) - \lk(x_j,x_i^+)$.
\end{proof}
\begin{cor}
If $k$ is a knot, then $\Delta_k(1)=1$, and if $L$ is a link, then $\Delta_L(1)=0$.
\end{cor}
\begin{proof}
Let $F_k$ and $F_L$ be Seifert surfaces for $k$ and $L$. By \cref{intersection forms and Seifert matrices} we have $\Delta_k(1)=\det Q_M=1$ and $\Delta_L(1)=\det Q_L = 0$.
\end{proof}
The Alexander polynomial of the connected sum of two knots has a very simple expression.
\begin{prop}
If $k_1$ and $k_2$ are knots, then $\Delta_{k_1 \# k_2}(t) = \Delta_{k_1}(t) \cdot \Delta_{k_2}(t)$.
\end{prop}
\begin{proof}
We can assume that $k_1$ and $k_2$ are situated in $S^3$ far apart, and that $F_1$ and $F_2$ are Seifert surfaces for $k_1$ and $k_2$ respectively that do not intersect. Then by taking the boundary sum $F_1 \sharp F_2$ we get a Seifert surface for $k_1 \# k_2$. If $S_1$ and $S_2$ are the Seifert matrices for $F_1$ and $F_2$ respectively, then the Seifert matrix of $F_1 \sharp F_2$ is
\[ \begin{pmatrix} S_1 & 0 \\ 0 & S_2 \end{pmatrix} \]
We clearly now have
\[ \Delta_{k_1\# k_2}(t) = \det(t^{1/2} S_1 - t^{-1/2} S_1^T) \cdot \det(t^{1/2} S_2 - t^{-1/2} S_2^T) = \Delta_{k_1}(t) \cdot \Delta_{k_2}(t) \]
\end{proof}
The Alexander polynomial has a clear relationship with the genus of a knot.
\begin{prop}
For a knot $k$ we have $\deg \Delta_k(t) \leq g(k)$.
\end{prop}
\begin{proof}
If $S$ is a Seifert matrix of order $2m$ for $k$, then $\deg \Delta_k(t) \leq m$, hence $\deg \Delta_k(t) \leq g(k)$.
\end{proof}
\begin{prop}
If $\det S \neq 0$ for some Seifert matrix $S$ of $k$, then $\deg \Delta_k(t) = g(k)$.
\end{prop}
\begin{proof}
If $S$ is a Seifert matrix associated to a Seifert surface of genus $g$, then we can write $\Delta_k(t) = t^{-g} \det(S-tS^T)$. The constant term of the polynomial $\det(S-tS^T)$ is just $\det S$, which is non-zero by assumption. Therefore the coefficient of $t^{-g}$ in $\Delta_k(t)$ is non-zero, and so $\deg \Delta_k(t) = g$.
\end{proof}
A link $L$ is said to be a \textbf{split link} if one can find an embedded 2-sphere in $S^3$ that separates some components of $L$ from the other components of $L$.
\begin{prop}
If $L$ is an oriented, split link $L$, then $\Delta_L(t) = 0$.
\end{prop}
\begin{proof}
Suppose $L = k_1 \cup k_2$ such that $k_1$ can be separated from $k_2$ by an embedded 2-sphere. Let $F_1$ and $F_2$ be Seifert surfaces for $k_1$ and $k_2$ respectively, and let $g_1$ and $g_2$ be the genera of $F_1$ and $F_2$ respectively. We can constructed a Seifert surface for $L$ by taking the connected sum $F_1 \# F_2$; that is, cut small, open discs out of $F_1$ and $F_2$ and glue in a tube $S^1 \times I$. Recall that a surface of genus $g$ with $n$ boundary components has first homology of rank $2g+n-1$, hence $H_1(F_1) \cong \mathbb Z^{2g_1}$ and $H_2(F_2) \cong \mathbb Z^{2g_2}$. However, $H_1(F_1 \# F_2) \cong \mathbb Z^{2g_1+2g_2+1}$, so this homology group has one extra generator in addition to the generators coming from $F_1$ and $F_2$. This generator can be represented by the meridian $\mu$ of the tube $S^1 \times I$. This generator does not link with the push offs of any of the generators from $F_1$ and $F_2$, hence there is an entire row and column of zeros in the Seifert matrix of $F_1 \# F_2$, and so $\Delta_L(t)=0$.
\end{proof}
\subsection{Knots in other 3-Manifolds}
\label{Knots in other 3-Manifolds}
Much of what we have said so far can be generalized to 3-manifolds other than $S^3$. For example, a knot $k$ in a 3-manifold $M$ is just a smoothly embedded copy of $S^1$ into $M$. Unfortunately, there is no analogous concept of diagram of knots in a general manifold, but any term we defined before without the use of a diagram can be defined for knots in a manifold. We will work towards extending other concepts and results to knots in a manifold, but almost always we are going to assume that the manifold is an integral homology sphere. These are manifolds $M$ with $H_*(M) = H_*(S^3)$, which is equivalent to simply $H_1(M)=0$. To see this look at the split exact sequence from the universal coefficient theorem
\[ 0 \longrightarrow \Ext(H_0(M),\mathbb Z) \longrightarrow H^1(M) \longrightarrow \Hom(H_1(M),\mathbb Z) \longrightarrow 0 \]
Since $H_1(M)=0$ we have the first map is an isomorphism, but $\Ext(\mathbb Z,\mathbb Z)=0$, hence $H^1(M)=0=H_2(M)$. Knots in integral homology spheres have the following property, which is analogous to a result we saw for true 3-spheres.
\begin{prop}
\label{knot complement in homology sphere is homology circle}
If $M$ is a homology sphere and $k$ an embedded knot, then $M\backslash k$ is a homology circle, that is $H_*(M\backslash k) = H_*(S^1)$.
\end{prop}
\begin{proof}
The proof is the same as in \cref{homology of knot complement in sphere}.
\end{proof}
Our first definition of linking number of two oriented, disjoint knots needed a knot diagram, so we cannot use this definition in a manifold. However, if $M$ is an integral homology sphere, and $k_1$ and $k_2$ knots in $M$, \cref{knot complement in homology sphere is homology circle} says that $H_1(M \backslash N(k_1)) = \mathbb Z$ is generated by $[\mu]$, where $\mu$ is any meridian of $N(k)$ (which is just the boundary of any normal cross-section). So, as we did before, we define the linking number $\lk(k_1,k_2)$ to be the integer such that $[k_2] = \lk(k_1,k_2) \cdot [\mu]$. We can also define the canonical longitude $\ell$ of a knot $k$ in $M$ to be the closed curve in $\partial(M \backslash N(k))$ such that $\lk(k,\ell)=0$.
Next we hope that Seifert surfaces exist in this more general context, and in fact they do.
\begin{lem}
Every knot $k$ in an integral homology sphere $M$ has a Seifert surface, i.e. a compact, orientable surface $F$ such that $\partial F=k$.
\end{lem}
\begin{proof}
\end{proof}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\comment{
Let $k$ be a knot in $M$ with canonical longitude $\ell$, and orient $k$ and $\ell$ such that the basis $(k,\ell,n)$ is the same as the orientation on $M$ ($M$ is orientable since it is an integral homology sphere), where $n$ is the inward point normal vector. The choice of meridian $\mu$ and longitude $\ell$ gives an identification $\varphi : \partial(M \backslash N(k)) \rightarrow S^1 \times S^1$. Let $p_0 : \partial(M \backslash N(k)) \rightarrow S^1$ be $p_0 = \pi_{S^1} \circ \varphi$, where $\pi_{S^1}$ is projection onto the second factor.
\begin{prop}
\label{extend projection to knot complement}
Let $K = M \backslash N(k)$ be the knot complement. The projection $p_0 : \partial K \rightarrow S^1$ extends to a map $p : K \rightarrow S^1$.
\end{prop}
\begin{proof}
Let $i : \partial K \rightarrow K$ be the inclusion. We claim that $p_0$ extends to $p$ if and only if the homotopy class $[p_0]$ is in the image of the induced map $i^* : [K,S^1] \rightarrow [\partial K,S^1]$. Clearly if such an extension exists then $[p_0] = i^*[p]$. Conversely, suppose there is some homotopy class of maps $[p]$ such that $[p_0] = i^*[p]$. This means that $p_0$ and $p \circ i$ are homotopic on $\partial K$, but we need them to be \emph{equal}. Let $H$ be a homotopy between $p \circ i$ and $p_0$ and let $U \subseteq M$ be a collar of $\partial K$ in $K$, i.e. $U$ is diffeomorphic to $\partial K \times [0,1]$ such that $\partial K \times \lcb 0 \rcb$ is mapped identically to $\partial K \subset U$. If we define $p' : K \rightarrow S^1$ to be $p$ outside $U$ and $H(x,t)$ on the part of $U$ corresponding to $(x,t) \in \partial K \times I$, then we see that $p'$ is an extension of $p_0$.
In \cref{Cohomology with Homotopy} we saw that we have a natural isomorphism of groups
\[ [\partial K,S^1] \cong H^1(\partial K) \]
since $S^1$ is a $K(\mathbb Z,1)$ space. The inclusion $i$ induces a map on cohomology, and by naturality of the above isomorphism, and naturality of the Poincar\'{e} isomorphisms we get the following commutative diagram from the long exact sequence of the pair $(K,\partial K)$
\[
\xymatrix
@R=3pc
@C=3pc
{
H_1(\partial K) \ar[r]^{i_*} & H_1(K) \\
H^1(\partial K) \ar[u]^{PD} \ar[r]_{\delta} & H^2(K,\partial K) \ar[u]_{PD}
}
\]
where $\delta$ is the connecting homomorphism. We have that $[\ell]$ is the Poincar\'{e} dual of $[p_0]$, so by commutativity of the above square we have $\delta[p_0] = PD^{-1}i_*[\ell] = 0$, since $[\ell]$ is the canonical longitude, hence $[p_0] \in \ker \delta$. But, by exactness we have $[p_0] \in \ker \delta = \image i^*$, hence $p_0$ extends to a map $p$ by our previous claim.
\end{proof}
This proposition can easily be extended to links. One might wonder if the map $p_0$ is a fibration of the knot complement over $S^1$. This is not always true, so it leads to the notion of a ``fibered'' knot, and will be discussed later. We can use \cref{extend projection to knot complement} to show that Seifert surfaces exist for links in integral homology spheres.
\begin{prop}
Given a link $L$ in an integral homology sphere $M$ there is a smoothly embedded, compact, orientable and connected surface $F$ such that $\partial F = L$.
\end{prop}
\begin{proof}
Let $p : K \rightarrow \partial D^2$ be the map constructed in \cref{extend projection to knot complement}, where $K$ is the link complement in $M$. We can slightly homotopy $p$ such that it is transverse to a point $* \in \partial D^2$, in which case $F' = p^{-1}(*)$ is a smoothly embedded submanifold of $K$ of codimension 1 (by \cref{pre-image of submanifold under transversality}) with boundary contained in $\partial K$, hence $F$ is a surface. We can connect the boundary of $F'$ to $L$ via annuli contained in the tubular neighborhoods of the link components to get a Seifert surface $F$.
\end{proof}
}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\newpage
\section{Handlebodies}
\label{Handlebodies}
\subsection{Handlebody Decompositions}
\label{Handlebody Decompositions}
Let $X$ be a smooth $n$-dimensional manifold with boundary. The process of attaching an $n$-dimensional $k$-handle to $X$ is to form the quotient space $X \cup_\varphi h$, where $h$ is a copy of $D^k \times D^{n-k}$, and where $\varphi : \partial D^k \times D^{n-k} \rightarrow \partial X$ is a smooth embedding. Notice that $\partial D^k \times D^{n-k}$ is only part of the boundary of $D^k \times D^{n-k}$. Also, the manifold $X \cup_\varphi h$ is not smooth, as it has corners. However, there is a canonical way to smooth the corners, so we may assume the quotient space is a smooth $n$-manifold with boundary. If two attaching maps $\varphi,\varphi' : \partial D^k \times D^{n-k} \rightarrow \partial X$ are isotopic, then the resulting manifolds $X \cup_\varphi h$ and $X \cup_{\varphi'} h$ are diffeomorphic. Also note that there is an obvious deformation retract of $D^k \times D^{n-k}$ onto $D^k \times 0$, hence $X \cup_\varphi h$ deformation retracts onto $X \cup_\varphi D^k \times 0$ (where $\varphi$ is suitably restricted in the second space), hence attaching a $k$-handle is the same as attaching a $k$-cell.
The anatomy of a handle is simple and the pieces have intuitive names. The map $\varphi$ is called the \textbf{attaching map}, while $\partial D^k \times D^{n-k}$ is called the \textbf{attaching region} (and sometimes the image of this set under $\varphi$). The subset $D^k \times 0$ is called the \text{core} of the handle, and $0 \times D^{n-k}$ the \textbf{cocore}. The region $\partial D^k \times 0$ the \textbf{attaching sphere}, and $0 \times \partial D^{n-k}$ the \textbf{belt sphere}. Finally, the number $k$ is called the \textbf{index} of the handle. See \cref{handle anatomy}.
\begin{figure}[tb]
\centering
\begin{tabular}{|l|l|}
\hline
\multicolumn{2}{|c|}{Handle Anatomy} \\
\hline
Attaching map & $\varphi : \partial D^k \times D^n \rightarrow \partial X$ \\
\hline
Attaching region & $\partial D^k \times D^{n-k}$ \\
\hline
Attaching sphere & $\partial D^k \times 0$ \\
\hline
Belt sphere & $0 \times \partial D^{n-k}$ \\
\hline
Core & $D^k \times 0$ \\
\hline
Cocore & $0 \times D^{n-k}$ \\
\hline
\end{tabular}
\ \linebreak
\includegraphics[scale=1.2]{graphics/handleanatomy}
\caption{Anatomy of an $n$-dimensional $k$-handle}
\label{handle anatomy}
\end{figure}
Unfortunately this picture is a little misleading since attaching a $k$-handle does not always look like attaching a ball along two disconnected pieces of its boundary. For example, attaching a 4-dimensional 2-handle means attaching a copy of $D^2 \times D^2$ along the solid torus $S^1 \times D^2$. The boundary of $D^2 \times D^2$ is a union of solid tori $S^1 \times D^2 \cup D^2 \times S^1$ that intersect at their boundaries $S^1 \times S^1$, and this decomposition of $S^3$ into two solid tori is clearly the genus 1 Heegaard splitting.
An embedding $\varphi : \partial D^k \times D^{n-k} \rightarrow \partial X$ is the same as embedding the trivial normal bundle of the attaching sphere $\varphi(\partial D^k \times 0)$ into $\partial X$. Therefore $\varphi$ determines, and is determined by, an embedding $\varphi_0 : \partial D^k \times 0 \rightarrow \partial X$ of the attaching sphere and an equivalence $f$ of the normal bundle $\nu \varphi_0(\partial D^k)$ with the trivial vector bundle $\partial D^k \times \mathbb R^{n-k}$; this is intuitively plausible, but is also the content of the Tubular Neighborhood Theorem. The embedding $\varphi_0$ is just a \textbf{knot} in $\partial X$, and the vector bundle equivalence $f$ is called a \textbf{(normal) framing} of $\partial D^k = S^{k-1}$. If the data $(\varphi_0,f)$ and $(\varphi_0',f')$ are smoothly isotopic, then the resulting manifolds $X \cup_\varphi h$ and $X \cup_{\varphi'} h$ are isotopic.
Fix a framing $f_0$ for $S^{k-1}$, i.e. a vector bundle equivalence
\[
\xymatrix
{
S^{k-1} \times \mathbb R^{n-k} \ar[r]^{f_0} \ar[d] & \nu \varphi_0(S^{k-1}) \ar[d] \\
S^{k-1} \ar[r]_{\varphi_0} & \varphi_0(S^{k-1})
}
\]
If $f$ is another framing of $S^{k-1}$, then $f^{-1} \circ f_0 : S^{k-1} \times \mathbb R^k \rightarrow S^{k-1} \times \mathbb R^{n-k}$ must map $(p,v)$ to $(p,g(p) \cdot v)$, where $g : S^{k-1} \rightarrow \GL(n-k)$ is smooth. Without loss of generality we can assume that $g(p) = I$ for a fixed base point $p \in S^{k-1}$. Therefore each framing determines an element of the group $\pi_{k-1}(GL(n-k))$, which is isomorphic to $\pi_{k-1}(O(n-k))$ by the deformation retract of $GL(n-k)$ onto $O(n-k)$, and it is even true that the set of framings are in one-to-one correspondence with elements of $\pi_{k-1}(O(n-k))$. However, the set of framings is not canonically isomorphic to $\pi_{k-1}(O(n-k))$, since the identification depends on $f_0$, but it is a principal homogeneous space for $\pi_{k-1}(O(n-k))$. Sometimes there is a canonical choice of $f_0$, and this will be useful later.
We can get a lot of information about attaching very low index handles and very high index handles. The attaching sphere for a $0$-handle $D^0 \times D^n$ is empty, so attaching a $0$-handle to $X$ is the same as simply taking the disjoint union of $X$ with $D^n$. The attaching sphere of a $1$-handle is a disjoint pair of points, and so if $\partial X$ is connected and non-empty there is a unique isotopy class of attaching maps (well, except in the case $n=2$ and $\partial X$ non-compact, in which case we can also interchange the two components of the attaching sphere). Further, since $\pi_0(O(n-1)) = \mathbb Z/2$ for $n \geq 2$ there are exactly two framings on $S^0$, hence there are exactly two manifolds one can obtain by attaching a $1$-handle to a $0$-handle (these are the $n$-dimensional analogs of annulus and \Mobius band).
Similarly, for $(n-1)$-handles, with $n \neq 2$, there is a unique framing of $S^{n-2}$ since $\pi_{n-2}(O(1)) = 0$. In the case $n=2$ we see that $(n-1)$-handles are just $1$-handles, which were discussed above. Attaching an $n$-handle is the same as gluing a copy of $D^n$ to $\partial X$ along $\partial D^n = S^{n-1}$. Since $\partial X$ is $(n-1)$-dimensional we see that we can only attach $D^n$ to boundary components of $X$ that are diffeomorphic to $S^{n-1}$. However, due to issues of exotic $n$-spheres and exotic diffeomorphisms of $n$-spheres we can only say that attaching an $n$-handle is unique in dimensions $\leq 6$.
The other combinations of $k$ and $n$, where $k$ is neither small nor close to $n$, are harder to understand. If $n-k > 2$, then there is only one isotopy class of embeddings $S^{k-1}$ into $(n-1)$-dimensional manifolds, hence all embeddings can be ``unknotted.'' So, the only thing that controls what we get from attaching a $k$-handle to a $0$-handle is the framing of an unknotted sphere in $S^{n-1}$, which are of course parameterized by $\pi_{k-1}(O(n-k))$. Suppose we are attaching a $k$-handle $h_2 = D^k \times D^{n-k}$ to a $0$-handle $h_1 = D^n$ such that the attaching sphere is the standardly embedded, unknotted sphere $S^{k-1} \subset S^{n-1}$. Without loss of generality we can assume that we have an identification $h_1 = D^k \times D^{n-k}$ and that the attaching map of $h_1$ is of the form $\varphi : \partial D^k \times D^{n-k} \rightarrow \partial D^k \times D^{n-k} \subset D^n$ such that $\pi_{\partial D^k} \circ \varphi$ is the identity, and the restriction of $\varphi$ to any slice $\lcb * \rcb \times D^{n-k}$ is an orthogonal linear map. Let $X$ be the manifold $h_1 \cup_\varphi h_2$. We claim that there is a canonical map $\pi : X \rightarrow S^k$ making $X$ into a $D^{n-k}$-bundle. Decompose $S^k$ as a union of two discs, the upper and lower hemispheres. Then $\pi$ projects the handle $h_1$ onto its its first factor, which is then identified with the upper hemisphere, and $\pi$ projects $h_2$ onto its first factor, which is then identified with the lower hemisphere. It is clear that this makes $X$ into a $D^{n-k}$-bundle, and in fact all $D^{n-k}$-bundles are of this form. Further, consider the framing $f_0$ corresponding to $\id_{\partial D^k \times D^{n-k}}$. This clearly gives the trivial bundle $X = S^k \times D^{n-k}$, and so this is a canonical element to associate to the identity in $O(n-k)$.
If we apply the above discussion to the case $k=2$, then we have that $D^{n-2}$-bundles over $S^2$ are classified by $\mathbb Z$ if $n=4$ and by $\mathbb Z/2$ if $n>4$. In this case we can get a canonical identification of these bundles with $\mathbb Z$ ($n=4$) and $\mathbb Z/2$ ($n>4$) by associating the trivial bundle with 0. For $n=4$, the integer associated to a $D^2$ bundle over $S^2$ is called the Euler number of the bundle. If $n > 4$, then we denote the non-trivial $D^{n-2}$ bundle over $S^2$ by $D^{n-2} \tilde\times S^2$, and we also define the associated $S^{n-3}$ bundle over $S^2$ by $S^{n-3} \tilde\times S^2 = \partial (D^{n-2} \tilde\times S^2)$.
Let $X$ be a compact $n$-manifold with boundary written as $\partial_+ X \sqcup \partial_- X$. If $X$ is oriented give $\partial_+ X$ the induced orientation and $\partial_- X$ the opposite of the induced orientation. A \textbf{handlebody decomposition} of $X$ is an identification of $X$ with an $n$-manifold obtained by attaching handles to $\partial_- X \times I$ such that $\partial_- X$ is identified with $\partial_- X \times 0$. A manifold with such a decomposition is called a \textbf{relative handlebody} built on $\partial_- X$, or if $\partial_- X = \emptyset$ then $X$ is simply called a \textbf{handlebody}.
The theory of handlebodies and Morse theory are closely related, so it is hard to discuss one without the other. Unfortunately we have now come to a point where we need to assume the reader has some knowledge of Morse theory. One can read this section concurrently with \cref{Morse Theory}.
According to the results in \cref{Morse Theory} we have that every smooth, compact $n$-manifold has a handlebody decomposition using Morse theory, and this can even be extended to non-compact manifolds with compact boundary. Amazingly every topological $n$-manifold with $n \neq 4$ also admits a handlebody decomposition, but this is not true in dimension $n=4$. For example, there is a 4-manifold with intersection form $E_8$ which does not admit a smooth structure or a handlebody decomposition.
\begin{prop}
\label{handlebodies with increasing index handles}
Any handlebody decomposition of a pair $(X,\partial_-X)$ can be modified (by isotoping attaching maps) so that handles are attached in order of increasing index, and handles of the same index can be attached in any order or simultaneously.
\end{prop}
\begin{proof}
Suppose we attach a $k$-handle $h_1$ and then an $\ell$-handle $h_2$ to a manifold $X$, with $\ell \leq k$. First we will perturb the attaching map of $h_2$ slightly so that its attaching sphere does not intersect $h_1$'s belt sphere (if needed). Note that $h_1$'s belt sphere, $0 \times \partial D^{n-k}$, is of dimension $n-k-1$ and $h_2$'s attaching sphere, $\partial D^\ell \times 0$, is of dimension $\ell-1$. The codimension of the intersection of generically positioned submanifolds is the sum of the submanifolds' codimension, so the codimension of $h_1$'s belt sphere intersected with $h_2$'s attaching sphere, in $\partial X$, is
\[ (n-1)-(n-k-1) + (n-1)-(\ell-1) = n-1-n+k+1+n-1-\ell+1 = n + k-\ell \geq n \]
Hence we can perturb the attaching sphere of $h_2$ slightly so that the spheres become entirely disjoint. Now take a smooth vector field on $X \cup h_1$ emanating from the cocore of $h_1$ radially, and flow the attaching sphere of $h_2$ along this vector field. This will push the attaching sphere of $h_2$ to $\partial X$, so we can first attach $h_2$ and then $h_1$.
\end{proof}
We will always assume that our handlebodies are ordered so that handles are attached with increasing index, and we will write $X_k$ to denote $\partial_-X \times I$ union all the handles of index $\leq k$.
A handlebody decomposition of a pair $(X,\partial_- X)$ naturally leads to a dual handlebody decomposition of $(X,\overline{\partial_+ X})$, where $k$-handles in the first decomposition become $(n-k)$-handles in the second decomposition. When we attach a $k$-handle $h$ to $X$, the points in $\partial D^k \times D^{n-k}$ become interior points of $X \cup h$ and the points in $D^k \times \partial D^{n-k}$ become boundary points in $X \cup h$. Other handles will be attached to the part $D^k \times \partial D^{n-k}$ of $h$, so we could reverse this and instead think of $h$ has an $(n-k)$-handle being attached to those other handles. Intuitively, we are turning the handlebody upside down. If the handlebody decomposition is given by a Morse function $f$, then the dual handlebody is just the handlebody given by the Morse function $1-f$.
\begin{example}
\label{handlebody decompositions of CP^n and RP^n}
We will show how $\mathbb CP^n$, complex projective $n$-space, can be decomposed into a handlebody with one $2k$-handle for each $0 \leq k \leq n$. Recall that $\mathbb CP^n$ is the quotient of $\mathbb C^{n+1}-0$ by $\mathbb C^*$ acting by scalar multiplication. The equivalence class of $(z_0,\ldots,z_n)$ is denoted by $[z_0:\cdots:z_n]$. Complex projective space is an $n$-dimensional complex manifold (or $2n$-dimensional real manifold) that can be covered with $n+1$ charts $\varphi_i : \mathbb C^n \rightarrow \mathbb CP^n$ defined by
\[ \varphi_i(z_0,\ldots,z_{n-1}) = [z_0:\cdots:z_{i-1}:1:z_i:\cdots:z_{n-1}] \]
The inverses $\phi_i : U_i \rightarrow \mathbb C^n$ of these charts are given by
\[ \phi_i([z_0,\ldots,z_n]) = (z_0/z_i,\ldots,z_{i-1}/z_i,z_{i+1}/z_i,\ldots,z_n/z_i) \]
where $U_i$ consists of all points $[z_0,\ldots,z_n]$ such that $z_i \neq 0$.