-
Notifications
You must be signed in to change notification settings - Fork 0
/
Copy paththesis.tex
2061 lines (1838 loc) · 104 KB
/
thesis.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
\documentclass[12pt,twoside]{reedthesis}
\usepackage{amsmath}
\usepackage{amssymb}
\usepackage{amsthm}
\usepackage{eucal}
\usepackage{tikz-cd}
\usepackage{mathtools}
\usepackage{booktabs}
\usepackage{graphicx}
\usepackage{braids}
\usepackage{color}
\usepackage[page]{appendix}
\usepackage[shortlabels]{enumitem}
\usepackage{minted}
\usepackage[noend]{algpseudocode}
\usepackage{hyperref}
\hypersetup{
colorlinks=true,
linkcolor=black,
citecolor=black,
urlcolor=black,
linkbordercolor=0 0 0,
menubordercolor=0 0 0
}
\usepackage{caption}
\usepackage{subcaption}
\captionsetup[subfigure]{format=hang}
\usepackage[backend=biber, style=alphabetic, sorting=anyt]{biblatex}
\addbibresource{thesis.bib}
\theoremstyle{definition}
\newtheorem{thm}{Theorem}[chapter]
\newtheorem{cor}[thm]{Corollary}
\newtheorem{conj}[thm]{Conjecture}
\newtheorem{lemma}[thm]{Lemma}
\newtheorem{defn}[thm]{Definition}
\newtheorem*{notation}{Notation}
\newtheorem{ex}[thm]{Example}
\newtheorem{notex}[thm]{Non-Example}
\newtheorem{prop}[thm]{Proposition}
\newtheorem{claim}[thm]{Claim}
\newtheorem*{note}{Note}
\newtheorem{facts}[thm]{Facts}
\newtheorem{fact}[thm]{Fact}
\usepackage{color}
\definecolor{strand}{RGB}{231, 76, 60}
% Use line breaks, not indents, for distinguishing paragraphs.
\setlength{\parindent}{0em}
\setlength{\parskip}{1em}
\setlist{topsep=0pt}
\newcommand{\N}{\mathbb{N}}
\newcommand{\Z}{\mathbb{Z}}
\newcommand{\Q}{\mathbb{Q}}
\newcommand{\R}{\mathbb{R}}
\newcommand{\C}{\mathbb{C}}
\renewcommand{\H}{\mathbb{H}}
\newcommand{\LS}{\mathcal{L}}
\newcommand{\SLZ}{\mathrm{SL}_2{\Z}}
\newcommand{\GLZ}{\mathrm{GL}_2{\Z}}
\newcommand{\SLR}{\mathrm{SL}_2{\R}}
\newcommand{\PSLZ}{\mathrm{PSL}_2{\Z}}
\newcommand{\PSLR}{\mathrm{PSL}_2{\R}}
\newcommand{\RP}{\mathrm{RP}^1}
\newcommand{\UT}{\mathrm{UT}}
\newcommand{\exptwothree}{\exp_{\{2,3\}}}
\renewcommand{\vec}[1]{\mathbf{#1}}
\DeclareMathOperator{\im}{im}
\DeclareMathOperator{\tr}{tr}
\DeclareMathOperator{\sign}{sign}
\newcommand{\into}{\xhookrightarrow{}}
\newcommand{\wo}{\, \backslash \,}
\definecolor{todopink}{RGB}{219, 48, 122}
\newcommand{\TODO}[1]{{\color{todopink}\textsf{TODO: #1}}}
\newcommand{\Id}{\mathrm{Id}}
\newcommand{\defnphrase}[1]{\textbf{#1}}
\DeclarePairedDelimiter\ang{\langle}{\rangle}
\DeclarePairedDelimiter\bigsawtooth{\Bigl( \! \! \Bigl(}{\Bigr) \! \! \Bigr)}
\DeclarePairedDelimiter\sawtooth{( \! (}{) \! )}
\newcommand{\mathe}[1]{\mintinline{Mathematica}|#1|}
\title{The Modular Flow in Three Homeomorphic Spaces}
\author{Christopher Henn}
\date{May 2018}
\division{Mathematics and Natural Sciences}
\advisor{Kyle M. Ormsby}
\department{Mathematics}
\begin{document}
% Fix a bug with display math vertical spacing. These commands must come after the start of the document.
\setlength{\abovedisplayshortskip}{1em}
\setlength{\belowdisplayshortskip}{1em}
\setlength{\abovedisplayskip}{1em}
\setlength{\belowdisplayskip}{1em}
\maketitle
\frontmatter
\pagestyle{empty}
\chapter*{Acknowledgements}
Many thanks,
\begin{itemize}
\item to Kyle, for his unceasing help in creating this thesis;
\item to my friends at O-Town, a cheery bunch, and to those at Reed;
\item and to my parents, for conceiving my existence and providing unconditional love and support.
\end{itemize}
I am deeply indebted to you all.
Thank you!
\setlength{\parskip}{0.2em}
\tableofcontents
\setlength{\parskip}{1em}
\chapter*{Abstract}
In this thesis we describe the modular flow, a dynamical system in a space of lattices $\LS_1$.
The modular flow is closely connected to a classically studied dynamical system---the geodesic flow on the modular surface---and presents in numerous and diverse fields of study.
We detail a recent knot-theoretic finding concerning the periodic orbits of the modular flow from the work of E. Ghys in 2007.
We then analyze periodic orbits of the modular flow in $\exptwothree S^1 \cong \LS_1$, a seldom-studied space consisting of all two- and three-point subsets of the circle $S^1$.
Using our analysis of the modular flow in $\exptwothree S^1$, we present a computer-aided diagrammatic system for decomposing homotopy-classes of periodic orbits as braids in the braid group $B_3$.
We finish with a conjecture relating these diagrams to the results of Ghys.
% \listoftables
% \listoffigures
\mainmatter
\pagestyle{fancyplain}
\chapter*{Introduction}
\addcontentsline{toc}{chapter}{Introduction}
\chaptermark{Introduction}
\markboth{Introduction}{Introduction}
The 3-sphere, the space of nontrivial lattices, and 1 to 3 point subsets of the circle---each of these spaces possesses a unique and rich structure.
Remarkably, they are also homeomorphic, providing multiple perspectives of phenomena occurring in each space individually.
In this document, we examine the modular flow and its periodic orbits in the space of nontrivial lattices.
This is a dynamical system whose periodic orbits can be described in terms of hyperbolic elements of the modular group, $\PSLZ$.
Viewed in the 3-sphere, such a flow presents as a link with the trefoil knot.
A basic property of these trefoil links, their linking number, can be described by a classic yet mysterious arithmetic function.
Moreover, the knot-complement of the trefoil in the 3-sphere has fundamental group isomorphic to a famous group, the braid group on 3 strands, providing an additional viewpoint from which to consider periodic orbits of lattices and their associated trefoil links.
The space of finite subsets of the circle is perhaps the most obscure of the aforementioned spaces, having been studied little in comparison to the space of lattices and the 3-sphere.
The modular flow in this finite subset space begs a visual interpretation similar to that of the trefoil links in the 3-sphere.
Here, we examine the group theoretic interpretation of finite subset loops as elements of the braid group.
Additionally, we provide a conjecture for determining the linking number of a trefoil link from such a loop.
This document will proceed by examining each homeomorphic space in turn.
\chapter{The Space of Lattices}
First, we recall some necessary definitions.
A \defnphrase{lattice in $\R^2$} is a discrete, closed, and additive subgroup of $\R^2$.
If a lattice $L$ is isomorphic to $\Z$ or $\Z \oplus \Z$, then we say that $L$ is a \defnphrase{degenerate} or \defnphrase{nondegenerate} lattice, respectively.
If $L$ is isomorphic to the trivial group, then $L$ is the \defnphrase{zero lattice}. A lattice of each type is shown in Figure~\ref{fig:three_types_of_lattices}.
We can construct any nondegenerate lattice using a real $2 \times 2$ matrix whose columns are linearly independent.
If $B$ is such a matrix, then
\begin{equation*}
B(\Z^2) = \{ B x : x \in \Z^2 \}
\end{equation*}
is the \defnphrase{lattice generated by $B$}.
We call $B$ a \defnphrase{basis} for the lattice and the columns of $B$ the \defnphrase{generators} of the lattice.
Degenerate lattices can be similarly constructed using a single generator.
Two lattices $L$ and $L'$ are \defnphrase{homothetic} if they are rescaled versions of each other, i.e. if $L' = tL = \{t x : x \in L\}$ for some fixed nonzero $t \in \R$.
We denote the space of all nontrivial lattices modulo the homothety relation as
\begin{equation*}
\LS = \LS_0 \cup \LS_1
\end{equation*}
where $\LS_0$ and $\LS_1$ are the degenerate and nondegenerate lattices, respectively.
\begin{figure}[h]
\centering
\begin{subfigure}[t]{0.31\textwidth}
\centering
\includegraphics[width=\textwidth]{figures/zero_lattice.pdf}
\caption{The zero lattice.}
\end{subfigure}
\hfill
\begin{subfigure}[t]{0.31\textwidth}
\centering
\includegraphics[width=\textwidth]{figures/degen_lattice.pdf}
\caption{A degenerate lattice.}
\end{subfigure}
\hfill
\begin{subfigure}[t]{0.31\textwidth}
\centering
\includegraphics[width=\textwidth]{figures/non_degen_lattice.pdf}
\caption{A nondegenerate lattice.}
\end{subfigure}
\caption{The three types of lattices.}
\label{fig:three_types_of_lattices}
\end{figure}
\section{The Modular Flow of Lattices}\label{subsec:lattice_flow}
We now describe a dynamical flow in the space of nondegenerate lattices.
Left multiplication of each point in a nondegenerate lattice by
\begin{equation}\label{eq:delta_t}
\delta_t = \begin{pmatrix}
\exp(t) & 0 \\
0 & \exp(-t)
\end{pmatrix}
\end{equation}
produces another nondegenerate lattice (where $t \in \R^+$).
As the time variable $t$ increases continuously, we obtain a flow in $\LS_1$ known as \defnphrase{the modular flow} \cite{ghys2007, touristguide, ncatcafe}.
Note that as $t$ increases, each individual point in the lattice moves towards $\pm \infty$ along a hyperbolic path, yet the lattice as a whole may move back to its original position.
Thus we can study the periodic orbits of this flow.
These periodic orbits can be described succinctly via a particular type of matrix.
Recall that that the \defnphrase{special linear group $\SLZ$} is the set of $2 \times 2$ matrices with integer entries and determinant 1.
The \defnphrase{projective special linear group $\PSLZ$} is $\SLZ / \{\pm I\}$.
We say that a matrix
\begin{equation*}
A = \begin{pmatrix}
a & b \\
c & d
\end{pmatrix} \in \PSLZ
\end{equation*}
is \defnphrase{hyperbolic} if $|a + d| > 2$.
Hyperbolic matrices are diagonalizable over the real numbers, so we can find a real $2 \times 2$ matrix $P$ such that
\begin{equation*}
PAP^{-1} = \begin{pmatrix}
\lambda_1 & 0 \\
0 & \lambda_2
\end{pmatrix}
\end{equation*}
where $\lambda_1$ and $\lambda_2$ are the eigenvalues of $A$.
Since $A = -A$ in $\PSLZ$, we can assume that $\lambda_1$ is positive.
Then since $\det(A) = 1$ and the product of eigenvalues of a matrix is equal to its determinant, we can find $P$ such that
\begin{equation*}
PAP^{-1} = \begin{pmatrix}
\exp(T) & 0 \\
0 & \exp(-T)
\end{pmatrix} = \delta_T
\end{equation*}
for some $T \in \R$.
Then
\begin{align*}
\delta_{T}P(\Z^2) &= PAP^{-1} P(\Z^2) \\
&= PA(\Z^2) \\
&= P(\Z^2) && \text{(see Lemma~\ref{lemma:m_is_integer}).}
\end{align*}
Thus the image of $P(\Z^2)$ by $\delta_t$ as $0 \leq t \leq T$ is a periodic orbit of orbit time $T$.
To summarize, we have the following proposition.
\begin{prop}\label{prop:hyperbolic_defines_flow}
Every hyperbolic element in $\PSLZ$ defines a periodic orbit of the modular flow.
\end{prop}
For a concrete example of these periodic orbits, consider the hyperbolic matrix
\begin{equation*}
A = \begin{pmatrix}
2 & 1 \\
1 & 1
\end{pmatrix}
\in \PSLZ
\end{equation*}
The eigenvalues of $A$ yield a diagonal matrix $D$ such that $PAP^{-1} = D$, where $P$ is the matrix whose columns are the eigenvectors of $A$.
Explicitly, we compute that
\begin{equation*}
D = \begin{pmatrix}
(3 + \sqrt{5}) / 2 & 0 \\
0 & (3 - \sqrt{5}) / 2
\end{pmatrix}, \quad P = \begin{pmatrix}
(1 + \sqrt{5}) / 2 & 1 \\
(1 - \sqrt{5}) / 2 & 1
\end{pmatrix}.
\end{equation*}
Let $\lambda$ be the first eigenvalue of $A$.
We know that $D = \delta_T$ for some $T$, hence $\lambda = e^T$ and $T \approx 0.96$ is the orbit time for the periodic orbit corresponding to $A$.
If we continuously act upon $P$ by $\delta_t$ as $0 \leq t \leq T$, we observe a periodic orbit in the space of lattices.
This orbit is visualized in Figure~\ref{fig:periodic_orbits_lattice}.
\begin{figure}[h]
\centering
\includegraphics[width=0.8\linewidth]{figures/periodic_orbits_lattice.pdf}
\caption{A periodic orbit of the modular flow in the space of lattices. As each point in the lattice given by $P$ is acted upon by $\delta_t$, it moves on a hyperbolic path towards infinity. Here, the path of each point is depicted by a sequence of points with shrinking radii. At $T \approx 0.96$, we have that $\delta_T P(\Z^2) = P(\Z^2)$.}
\label{fig:periodic_orbits_lattice}
\end{figure}
It is convenient to label a periodic orbit of the modular flow by the matrix $A \in \PSLZ$ that produced it.
But does every periodic orbit arise from an element of $\PSLZ$?
The following results establish that this is indeed the case.
\begin{lemma}\label{lemma:m_is_integer}
Matrices $X$ and $Y$ in $\PSLR$ generate the same lattice if and only if $X = YA$ for some $A \in \PSLZ$.
In particular, if $X$ generates the square lattice, then $X \in \PSLZ$.
\end{lemma}
\begin{proof}
Suppose that $X$ and $Y$ generate the same lattice.
Since a lattice consists of all integer linear combinations of its generators, we know that there exists some integer matrix $A$ such that $X = YA$.
Because $X$ and $Y$ are in $\PSLR$, we also know that $\det(X) = \det(Y) = 1$. Then
\begin{equation*}
1 = \det(X) = \det(YA) = \det(Y)\det(A) = 1 \cdot \det(A)
\end{equation*}
and so $\det(A) = 1$ and $A \in \PSLZ$.
To show the converse, suppose instead that $X = YA$ for some $A \in \PSLZ$.
Then each column of $X$ is in $Y(\Z^2)$, thus $X(\Z^2) \subseteq Y(\Z^2)$.
In addition, $Y = XA^{-1}$ and $A^{-1} \in \PSLZ$, so similarly we have that $Y(\Z^2) \subseteq X(\Z^2)$.
If $X$ generates the square lattice, then $X(\Z^2) = I(\Z^2)$, and so $X = IA = A \in \PSLZ$.
\end{proof}
\begin{prop}
The periodic orbits of the modular flow are in bijection with conjugacy classes of elements of $\PSLZ$.
\end{prop}
\begin{proof}
Given a matrix $A \in \PSLZ$, we have already shown how to construct a periodic orbit of the modular flow.
It remains to be shown that any element in the conjugacy class of $A$ produces the same periodic orbit, and that any periodic orbit of the modular flow defines the conjugacy class of some hyperbolic element in $\PSLZ$.
Let $P \in \PSLR$ such that $PAP^{-1} = \delta_t$ for some $t \in \R$, so that $\delta_t P(\Z^2) = P(\Z^2)$.
Also suppose that $B \in \PSLZ$ as well, so that $BAB^{-1}$ is in the conjugacy class of $A$.
Then
\begin{equation*}
(PB^{-1}) BAB^{-1} (PB^{-1})^{-1} = PAP^{-1} = \delta_t
\end{equation*}
but also
\begin{equation*}
(PB^{-1})(\Z^2) = P(B^{-1} (\Z^2)) = P(\Z^2).
\end{equation*}
Hence the element $BAB^{-1} \in \PSLZ$ defines a periodic orbit with the same orbit time and same initial lattice $P(\Z^2)$, i.e. the same periodic orbit as $A$.
Now, suppose instead that we are given a periodic orbit of the modular flow.
We wish to show that it corresponds to the conjugacy class of some hyperbolic element in $\PSLZ$.
The periodic orbit can be described by some $M \in \PSLR$ and $t \in \R$ where $\delta_t M(\Z^2) = M(\Z^2)$
Observe that if $A = M^{-1} \delta_t M$, then
\begin{equation*}
A(\Z^2) = M^{-1} \delta_t M (\Z^2) = M^{-1} M (\Z^2) = \Z^2.
\end{equation*}
Thus by Lemma~\ref{lemma:m_is_integer}, $A \in \PSLZ$.
Note, however, that our choice of $M$ was not unique.
Let $N \in \PSLR$ be such that $M(\Z^2) = N(\Z^2)$.
Then again by the lemma, we have that $N = MB$ for some $B \in \PSLZ$, so
\begin{equation*}
N^{-1} \delta_t N = (MB)^{-1} \delta_t (MB) = B^{-1} M^{-1} \delta_t M B = B^{-1} A B.
\end{equation*}
Hence every periodic orbit defines a conjugacy class of some element $A \in \PSLZ$.
We now show that $A$ is hyperbolic by computing its trace directly.
Since $MAM^{-1} = \delta_t$ and trace is invariant under conjugation, we have that
\begin{align*}
\tr(A) &= \tr(MAM^{-1}) \\
&= \tr(\delta_t) \\
&= e^{t} + e^{-t} \\
&= \left( \sum_{k=0}^\infty \frac{t^k}{k!} \right) + \left( \sum_{k=0}^\infty \frac{(-t)^k}{k!} \right) \\[0.5em]
&= \sum_{k=0}^\infty \frac{t^k + (-t)^k}{k!} \\[0.5em]
&= 2 + 0 + \frac{t^2 + t^2}{2!} + 0 + \frac{t^4 + t^4}{4!} + \cdots \\
&> 2.
\end{align*}
Thus $A$ is indeed hyperbolic. To summarize, there are two maps
\begin{equation*}
\left\lbrace
\begin{array}{c}
\text{conjugacy classes of} \\
\text{hyperbolic } A \in \PSLZ
\end{array}
\right\rbrace \rightleftarrows \left\lbrace
\begin{array}{c}
\text{periodic orbits of} \\
\text{the modular flow}
\end{array}
\right\rbrace
\end{equation*}
and it is straightforward to observe that they form a bijection.
\end{proof}
\section{The Classical Perspective on the Modular Flow}
Our construction of the modular flow is a non-standard take on a well-known and classically studied dynamical system called \emph{the geodesic flow on the modular surface}.
Here, we provide a brief sketch of the more classical construction of this system and then motivate the alternative perspective we have just detailed.
\begin{figure}[b!]
\centering
\includegraphics[width=0.8\linewidth]{figures/tessellation.pdf}
\caption{The elements of $\mathcal{M}$ are orbits of $\mathbb{H}$ under the action of $\PSLZ$. Every orbit has exactly one representative in the shaded region $\mathcal{F}$. The image of $\mathcal{F}$ by various elements of the modular group (in terms of the modular group generators $S$ and $T$) are shown alongside $\mathcal{F}$. These translates are called \defnphrase{faces} in a \defnphrase{tessellation} of $\mathbb{H}$ by its fundamental domain.}
\label{fig:tessellation}
\end{figure}
Let $\mathbb{H} = \{ x + i y \in \C : y > 0,\ x,\, y \in \R \}$ be the upper half plane.
There is an action of the \defnphrase{modular group} $\PSLZ = \SLZ / \{ \pm I \}$ on $\mathbb{H}$ defined by
\begin{align*}
\PSLZ \times \mathbb{H} &\to \mathbb{H} \\
(\gamma, \tau) &\mapsto \gamma\tau = \frac{a \tau + b}{c \tau + d}
\end{align*}
where
\begin{equation*}
\gamma = \begin{pmatrix}
a & b \\
c & d
\end{pmatrix} \in \PSLZ.
\end{equation*}
Such a map is called a \defnphrase{linear fractional transform}.
The \defnphrase{modular surface $\mathcal{M}$} is the quotient of $\mathbb{H}$ by this action of $\PSLZ$.
More explicitly, $\mathcal{M}$ is the set of all orbits $\{ \gamma \tau : \gamma \in \PSLZ \}$ as $\tau$ varies over of $\mathbb{H}$.
The subset
\begin{equation*}
\mathcal{F} = \{ \tau \in \mathbb{H} : |\mathrm{Re}(x)| \leq 1/2,\ |\tau| \geq 1 \}
\end{equation*}
of $\mathbb{H}$ contains exactly one point from each one of these orbits (except on its boundary), and thus provides a convenient geometrical region that represents the modular surface (the region $\mathcal{F}$ is a \defnphrase{fundamental domain} for $\mathcal{M}$).
The right and left vertical boundary of $\mathcal{F}$ are identified by the action of $\PSLZ$, as well as the right and left bottom circular boundary.
In Figure~\ref{fig:tessellation}, we visualize $\mathcal{F}$ along with its image under several elements of $\PSLZ$.
For a more thorough treatment of $\mathbb{H}$, $\PSLZ$, and $\mathcal{F}$, see \cite{katok1992}.
Recall that an \defnphrase{$n$-dimensional manifold} is a generalization of $n$-dimensional Euclidean space; some small neighborhood of every point on a manifold looks like $n$-dimensional Euclidean space.
A \defnphrase{Riemannian manifold} is a smooth and real manifold $M$ equipped with an Riemannian inner product on the tangent space $T_pM$ for each point $p$ in $M$.
A \defnphrase{geodesic path} is the generalization of a straight line in Euclidean space to Riemannian manifolds---the quickest way between two points across the surface of a manifold is via a geodesic path, just as the quickest way between two points in Euclidean space is via a straight line.
The prototypical example of a geodesic is a curved arc across the surface of the 2-sphere.
In $\H$, geodesics under the Poincar\'e metric
\begin{equation*}
ds = \frac{\sqrt{dx^2 + dy^2}}{y},
\end{equation*}
consist of straight vertical lines originating at the $x$-axis, and half-circles that meet the $x$-axis at right angles (see Figure~\ref{fig:geodesics_in_h}).
\begin{figure}[t]
\centering
\includegraphics[width=0.6\linewidth]{figures/geodesics_in_h.pdf}
\caption{Example geodesics in $\mathbb{H}$.}
\label{fig:geodesics_in_h}
\end{figure}
Now, suppose that $M$ is a Riemannian manifold.
Given a point $p \in M$ and a direction $v$ in the space tangent to $M$ at $p$, the \defnphrase{geodesic flow on $M$} through $p$ in the direction of $v$ can be simply described as the flow along the geodesic in $M$ that points in the direction of $v$.
A \defnphrase{closed orbit of the geodesic flow} is a geodesic that returns to its initial position pointing in the same initial direction $v$.
The physical interpretation of the geodesic flow is to think of a particle moving on a manifold $M$ subject to no constraints other than that it must stay on $M$.
Many fundamental equations in physics, such as the Euler equations of motion of a rigid body or the Euler equations of fluid dynamics of an inviscid incompressible fluid, are examples of the geodesic flow on particular manifolds \cite{arnold66, tao}.
For Riemannian manifolds, one may also define the unit tangent bundle.
The \defnphrase{unit tangent bundle $\UT(M)$ of a manifold $M$} consists of pairs $(p, v)$, where $p \in M$ and $v$ is a unit tangent vector to $M$ at $p$.
There exists a natural projection $\pi : (p, v) \mapsto p$ mapping $UT(M)$ onto $M$.
The space of nontrivial lattices up to homothety $\LS_1$ is isomorphic to the unit tangent bundle for the modular surface $\mathcal{M}$.
Thus every lattice $L \in \LS_1$ implicitly describes a point in $\mathcal{M}$ and a unit tangent direction---all the data needed to define the geodesic flow on $\mathcal{M}$.
When we study the ``modular flow'', we are really studying the lift of the geodesic flow on $\mathcal{M}$ into the tangent space $\mathrm{UT}(M)$.
If we project an orbit of the modular flow back down into $\mathcal{M}$, we recover a (rescaled) orbit of the geodesic flow on $\mathcal{M}$.
Suppose $L \in \LS_1$ is a lattice generated by $\omega_1$ and $\omega_2$ such that $\omega_1,\, \omega_2 \in \H$ and the angle between $\omega_1$ and $\omega_2$ is positive and less than 180 degrees.
Then the map $L \mapsto \omega_1 / \omega_2$ is a projection from $\LS_1 = \UT(\mathcal{M})$ to $\mathcal{M}$ (see \cite[6-13]{silverman1994}).
The projection onto $\mathcal{M}$ of several orbits of the modular flow in $\LS_1$ are visualized in Figure~\ref{fig:modular_surface_orbits}.
\begin{figure}[t]
\centering
\begin{subfigure}[t]{0.4\textwidth}
\centering
\includegraphics[width=\textwidth]{figures/modular_surface_orbit_pppqq.pdf}
\caption*{$A = \begin{pmatrix}7 & 3 \\ 2 & 1\end{pmatrix}$}
\end{subfigure}
\hspace{3mm}
\begin{subfigure}[t]{0.4\textwidth}
\centering
\includegraphics[width=\textwidth]{figures/modular_surface_orbit_ppppq.pdf}
\caption*{$A = \begin{pmatrix}5 & 4 \\ 1 & 1\end{pmatrix}$}
\end{subfigure}
\caption{Periodic orbits of modular flow corresponding to matrices $A \in \PSLZ$, projected down into $\mathcal{M}$ and visualized in the fundamental domain $\mathcal{F}$. In $\mathcal{F}$, the right and left vertical boundaries are identified, and the right and left bottom circular boundaries are identified.}
\label{fig:modular_surface_orbits}
\end{figure}
The idea to consider a lift of the geodesic flow on $\mathcal{M}$ rather than the geodesic flow on $\mathcal{M}$ itself is a recent and novel contribution from Ghys in \cite{ghys2007}.
In addition to being homeomorphic to $\LS_1$, the unit tangent bundle of $\mathcal{M}$ is homeomorphic to the knot complement of the trefoil knot $S^3 \wo K$, as we shall see in Section~\ref{sec:lattice_to_s3}.
Thus Ghys studies the periodic orbits of the geodesic flow \emph{as knots}, and obtains an interesting result relating a simple algebraic invariant of these knots back to the modular group $\PSLZ$.
We summarize this result in Section~\ref{sec:periodic_orbits_in_s3}.
A third space, $\exp_{\{2,3\}} S^1$, is homeomorphic to the unit tangent bundle of $\mathcal{M}$ as well.
In Chapter~\ref{chap:finite_subsets}, we use this third space to analyze homotopy classes of periodic orbits in each space homeomorphic to the unit tangent bundle.
To summarize, the rest of this document is concerned with the modular flow in the upper row of the following diagram:
\begin{center}
\begin{tikzcd}
\exp_{\{2,3\}} S^1 \ar[r, "\cong", leftrightarrow] & \LS_1 \ar[d, "\pi"] & S^3 \wo K \ar[l, "\cong"', leftrightarrow] \\
& \mathcal{M} &
\end{tikzcd}
\end{center}
\chapter{The 3-sphere and the Trefoil Complement}\label{sec:three_sphere}
\section{The Fundamental Group of the Trefoil Complement}
The \defnphrase{3-sphere} is the set of complex tuples
\begin{equation*}
S^3 = \{ (u, v) \in \C : |u|^2 + |v|^2 = 1 \}
\end{equation*}
This set is endowed with the usual subspace topology, and is commonly employed in the study of knots or links.
Technically speaking, we say that a \defnphrase{knot} is an embedding $S^1 \into S^3$, considered up to isotopy.
Intuitively, a knot is a loop in 3-space with interesting twisting behavior and no self-intersections.
\begin{figure}[h]
\centering
\includegraphics[width=0.5\linewidth]{figures/trefoil.png}
\caption{A trefoil knot. While the knot is embedded inside $S^3$, we can stereographically project $S^3$ to $\R^3$ for visualization purposes.}
\label{fig:trefoil}
\end{figure}
In our case, we're interested in a particular prototypical knot called the \defnphrase{trefoil} (visualized in Figure~\ref{fig:trefoil}).
This is the knot that, when laid on the surface of the torus, crosses the meridian of the torus three times and crosses longitudinal circles twice---as such it is also known as the \defnphrase{$(2, 3)$-torus knot}.
For any knot $K : S^1 \into S^3$, we can study the \defnphrase{knot complement} space $S^3 \wo K$.
Historically, such spaces have provided useful knot invariants via their fundamental group and interesting examples of 3-manifolds.
The fundamental group of the trefoil complement happens to be isomorphic to a famous group with a convenient physical interpretation, which we will describe shortly.
First, we recall the definition of a fundamental group as well as a useful theorem for computing fundamental groups.
\begin{defn}
Let $X$ be a topological space. A \defnphrase{loop in $X$ based at $p$} is a continuous map $f : I \to X$ such that $f(0) = f(1) = p$.
Two loops $f$ and $g$ based at $p$ are \defnphrase{path-homotopic} if there exists a continuous map $H : I \times I \to X$ such that for all $s$ and $t$ in $I$, $H(s, 0) = f(s)$, $H(s, 1) = g(s)$, and $H(0, t) = H(1, t) = p$.
The path-homotopy relation is an equivalence relation on the set of all loops in $X$ based at $p$, and we call the equivalence class of a loop its \defnphrase{path class}.
The \defnphrase{fundamental group of $X$ based at $p$} is the set of path classes of loops in $X$ based at $p$, denoted $\pi_1(X, p)$.
The set $\pi_1(X, p)$ forms a group under the operation of path composition.
\end{defn}
Morally, two loops in $X$ are homotopic if they can be continuously deformed into one another.
The fundamental group describes the group structure of all based loops in $X$ up to homotopy.
There is no general method for computing the fundamental group of an arbitrary topological space.
Instead, we rely on a toolkit of methods and theorems that prove useful in different situations.
One essential such tool is the Seifart Van-Kampen theorem.
\begin{defn}
Let $H$, $G_1$, and $G_2$ be groups with presentations
\begin{align*}
G_1 &\cong \ang{\alpha_1, \dots, \alpha_m : \rho_1, \dots, \rho_r}, \\
G_2 &\cong \ang{\beta_1, \ldots, \beta_n : \sigma_1, \ldots, \sigma_s}, \text{ and } \\
H &\cong \ang{\gamma_1, \ldots, \gamma_p : \tau_1, \ldots, \tau_t}.
\end{align*}
Then if $f_1 : H \to G_1$ and $f_2 : H \to G_2$ are group homomorphisms, the \defnphrase{amalgamated free product} of $G_1$ and $G_2$ along $H$ is the group with presentation
\begin{align*}
G_1 \ast_H G_2 \cong \ang{&\alpha_1, \ldots \alpha_m,\, \beta_1, \ldots, \beta_n : \rho_1, \ldots, \rho_r, \\
&\sigma_1, \ldots, \sigma_s,\, u_1 = v_1,\ldots, u_p = v_p}
\end{align*}
where $u_a$ is an expression for $f_1(\gamma_a) \in G_1$ in terms of the generators $\{\alpha_i\}$ and $v_a$ similarly expresses $f_2(\gamma_a) \in G_2$ in terms of $\{\beta_j\}$.
\end{defn}
\begin{thm}[Seifert-van Kampen Theorem]
Let $X$ be a topological space, and suppose that $X = U_1 \cup U_2$ for two open and path-connected subsets $U_1,\ U_2 \subseteq X$.
In addition, suppose that $U_1 \cap U_2$ is path-connected and nonempty, and let $p \in U_1 \cap U_2$.
There exists an isomorphism
\begin{equation*}
\Phi : \pi_1(U_1, p) \ast_{\pi_1(U_1 \cap U_2,\, p)} \pi_2(U_2, p) \to \pi_1(X, p)
\end{equation*}
such that the following diagram commutes:
\begin{center}
\begin{tikzcd}
& \pi_1(U_1) \ar[dr, bend left=20, "j_1",] \ar[d, hookrightarrow] & \\
\pi_1(U_1 \cap U_2) \ar[ur, bend left=20, "i_1"] \ar[dr, bend right=20, "i_2"'] & \pi_1(U_1) \ast_{\pi_1(U_1 \cap U_2)} \pi_2(U_2) \ar[r, "\Phi", dashed] \ar[d, hookleftarrow] & \pi_1(X) \\
& \pi_1(U_2) \ar[ur, bend right=20, "j_2"'] &
\end{tikzcd}
\end{center}
\end{thm}
Now, we provide an explicit computation of the fundamental group of the trefoil complement.
For additional details, see~\cite[47--49]{hatcher2002} and~\cite[153--154]{stillwell1993}.
\begin{prop}
Let $K$ be the trefoil knot.
Then
\begin{equation*}
\pi_1(S^3 \wo K) = \ang{x, y : x^2 = y^3}
\end{equation*}
\end{prop}
\begin{proof}
We proceed by two applications of the Seifert-van Kampen theorem.
First we compute $\pi_1(\R^3 \wo K)$, since it is slightly easier to compute than $\pi_1(S^3 \wo K)$.
Let $T \subset \R^3 \wo K$ be the solid torus with $K$ removed and $N$ be a thin tubular neighborhood of $K$.
Let $X = T \wo N$ and $Y = (\R^3 \wo T) \wo N$.
Then $X$ and $Y$ meet at a region $L$ that looks like the surface of the solid torus with $N$ removed.
As shown in Figure~\ref{fig:flat_annulus}, $L$ is an annulus.
\begin{figure}[h]
\centering
\includegraphics[width=0.7\textwidth]{figures/fundamental_group_flat.pdf}
\caption{The region $L$ is the torus with a thin tubular neighborhood of the trefoil $N$ removed. Here, $N$ is depicted in white and $L$ is shaded in gray on the flat torus. While still respecting the appropriate edge identifications, the pieces of the torus can be rearranged to form an annulus. The generator $\ell$ of $\pi_1$ of this annulus is shown in blue.}
\label{fig:flat_annulus}
\end{figure}
Since $L$ is an annulus, its fundamental group is infinite cyclic with a single generator $\ell$.
The space $X$ is a solid torus with a neighborhood of the trefoil removed, which deformation retracts onto a circle.
So $\pi_1(X)$ is infinite cyclic as well, with generator $x$ corresponding to the loop around the axis of the solid torus.
In $X$, the generator $\ell$ is a trefoil loop around the torus that crosses longitudinal circles twice, so we have
\begin{equation*}
\ell \simeq x^2.
\end{equation*}
The fundamental group of $Y$ is also infinite cyclic, with the generator $y$ corresponding to a loop that ``links'' the hole of the solid torus.
The loop $\ell$ crosses meridian circles in the solid torus three times, so we have that
\begin{equation*}
\ell \simeq y^3.
\end{equation*}
Now, we can expand $X$ and $Y$ slightly to open sets $X' \supseteq X$ and $Y' \supseteq Y$ such that $X' \cap Y'$ is an open neighborhood of $L$ and $X' \cup Y' = \R^3 \wo K$ (see Figure~\ref{fig:fundamental_group_side}).
Then, applying Van Kampen, we have that
\begin{equation*}
\pi_1(\R^3 \wo K) = \ang{x, y : x^2 = y^3}.
\end{equation*}
\begin{figure}[t]
\centering
\includegraphics[width=0.6\textwidth]{figures/fundamental_group_side.pdf}
\caption{The picture on the left shows a longitudinal slice of the torus $T$, with the absence of tubular neighborhood $N$ of the trefoil shown in white. The regions $X$ and $Y$ are depicted in blue and pink, respectively. On the right, $X$ and $Y$ have been expanded to $X'$ and $Y'$, respectively.}
\label{fig:fundamental_group_side}
\end{figure}
Now we show that $\pi_1(\R^3 \wo K) = \pi_1(S^3 \wo K)$ with another application of Van Kampen.
Decompose $S^3 \wo K$ into $(\R^3 \wo K) \cup B$, where $B$ is an open set containing the compactification point and the complement of a large closed ball containing $K$.
Both $B$ and $B \cap (\R^3 \wo K)$ are simply connected.
By Van Kampen, the inclusion $\R^3 \wo K \into S^3 - K$ induces an isomorphism of groups
\begin{equation*}
\Phi : \pi_1(R^3 \wo K,) \to \pi_1(S^3 \wo K)
\end{equation*}
as shown by the following commutative diagram:
\begin{center}
\begin{tikzcd}
& \pi_1(\R^3 \wo K) \ar[dr, bend left=20] \ar[d, hookrightarrow] & \\
\pi_1(B \cap (\R^3 \wo K)) \ar[ur, bend left=20] \ar[dr, bend right=20, "1"'] & \pi_1(B) \ast \pi_1(\R^3 \wo K) \ar[r, "\Phi", dashed] \ar[d, hookleftarrow] & \pi_1(S^3 \wo K) \\
& \pi_1(B) \ar[ur, bend right=20, "1"'] &
\end{tikzcd}
\end{center}
Thus we have the desired result.
\end{proof}
The fundamental group of the trefoil complement is isomorphic to a well-known group, the \defnphrase{braid group on $3$ strands}.
This group has presentation
\begin{equation*}
B_3 = \ang{\sigma_1, \sigma_2 : \sigma_1 \sigma_2 \sigma_1 = \sigma_2 \sigma_1 \sigma_2}.
\end{equation*}
The isomorphism $\pi_1(S^3 \wo K) \to B_3$ is given by sending generators
\begin{equation}\label{eq:iso_pi1_b3}
\begin{aligned}
x &\mapsto \sigma_0 \sigma_1 \sigma_0 \\
y &\mapsto \sigma_0 \sigma_1
\end{aligned}
\end{equation}
This group appears in many different contexts: it is the mapping class group of the thrice-punctured disk, the universal central extension of $\PSLZ$, and so forth.
It also has a convenient visual interpretation, where the generators $\sigma_1$ and $\sigma_2$ are depicted and composed as in Figure~\ref{fig:visual_braid_group}.
We will return to this visual interpretation in Chapter~\ref{chap:finite_subsets}, where we decompose the path class of various loops in $\exptwothree S^1 \cong S^3 \wo K$ as visual generators of $B_3$.
\vspace{1em}
\begin{figure}[h]
\centering
\begin{subfigure}[t]{0.48\textwidth}
\centering
\begin{tikzpicture}
\braid[ultra thick, number of strands=3, width=5mm, height=21mm] s_1;
\end{tikzpicture}
\hspace{10mm}
\begin{tikzpicture}
\braid[ultra thick, number of strands=3, width=5mm, height=21mm] s_2;
\end{tikzpicture}
\caption{The generators $\sigma_1$ and $\sigma_2$ of $B_3$}
\end{subfigure}
\hfill
\begin{subfigure}[t]{0.48\textwidth}
\centering
\begin{tikzpicture}
\braid[ultra thick, width=5mm, height=7mm, style strands={2}{strand}] s_1 s_2 s_1;
\draw node[fill=white] at (2,-1.25) {=};
\end{tikzpicture}
\begin{tikzpicture}
\braid[ultra thick, width=5mm, height=7mm, style strands={2}{strand}] s_2 s_1 s_2;
\end{tikzpicture}
\caption{The single relation in $B_3$, known as the \textbf{third Reidmeister move}.}
\end{subfigure}
\caption{The visual interpretation of $B_3$, where each generator is thought of as a braiding of three strands of string that start and end at fixed points. The group operation corresponds to stacking the visual generators top to bottom to compose a more complex braiding of string. The single relation in $B_3$ can be thought of as ``pulling a string'' without moving the endpoints, which preserves the isotopy class of the braid.}
\label{fig:visual_braid_group}
\end{figure}
\section{From $\LS$ to $S^3$}\label{sec:lattice_to_s3}
At first glance, the space of two-dimensional lattices up to homothety and the three-dimensional sphere may seem quite different.
It turns out that they are indeed homeomorphic.
We now describe an explicit homeomorphism $\LS \to S^3$ as sketched by Mostovoy in~\cite{mostovoy2004}.
Let $L \in \LS$ and $k \in \Z$.
Define
\begin{equation*}
G_k(L) = \sum_{\omega \in L,\ \omega \neq 0} \omega^{-k}.
\end{equation*}
When $k \geq 4$, this series converges and is known as an \defnphrase{Eisenstein series}.
For nondegenerate $L$, these numbers are the parameters in the Weierstrass $\wp_L$-function identity
\begin{equation*}
(\wp_L')^2 = 4 \wp_L^3 - 60 G_4(L) \wp_L - 140 G_6(L).
\end{equation*}
The following facts about $G_4(L)$ and $G_6(L)$ are fundamental to the theory of elliptic functions:
\begin{enumerate}[(i)]
\item For $t > 0$ and any $L$, $G_4(tL) = t^{-4} G_4(L)$ and $G_6(tL) = t^{-6} G_6(L)$~\cite[631]{abramowitz1964}.
\item The map $L \mapsto (G_4(L),\, G_6(L))$ is a homeomorphism $\LS \to \C^2 \wo \{ 0 \}$~\cite[82, 89]{serre1973}.
\item A tuple $(u, v) \in \C^2 \wo \{ 0 \}$ is the image of a degenerate lattice under this map if and only if $20u^3 - 49v^2 = 0$~\cite[265]{ghys2007}.
\end{enumerate}
Thus for any nontrivial lattice $L$, we can obtain a unique rescaling $L' = tL$ for some $t > 0$ such that
\begin{equation*}
|G_4(L')|^2 + |G_6(L')|^2 = 1
\end{equation*}
Then the map $L \mapsto (G_4(L'),\, G_6(L'))$ lands in $S^3$, and we have the desired homeomorphism $\LS \to S^3$.
The image of the degenerate lattices under this map is the intersection of the curve $20u^3 - 49v^2 = 0$ with the 3-sphere. For any $a \in \R$, it happens that the intersection of complex curve $u^3 + av^2 = 0$ with the 3-sphere is a trefoil knot \cite[4]{milnor1968}. Thus, the image of degenerate lattices under the homeomorphism is the trefoil shown in Figure~\ref{fig:trefoil}.
\section{The Linking Number of Periodic Orbits in $S^3$}\label{sec:periodic_orbits_in_s3}
We have observed the periodic orbits of the modular flow in the space of lattices in Section~\ref{subsec:lattice_flow}, and we have just seen that the space of lattices is homeomorphic to the 3-sphere.
It is now natural to ask: how do these periodic orbits present in the 3-sphere?
The degenerate lattices are sent to the trefoil knot in the 3-sphere, but the periodic orbits of this flow exist entirely within the space of nondegenerate lattices.
Thus we can say that in the 3-sphere, a periodic orbit of the modular flow in the space of lattices determines a link with the trefoil knot.
Since each periodic flow is associated with some hyperbolic $A \in \PSLZ$ we call the corresponding trefoil link $k_A$ a \defnphrase{modular trefoil link} or \defnphrase{modular knot}, following after~\cite{ghys2007}.
Several examples of such links are shown in Figure~\ref{fig:trefoil_links}.
\begin{figure}[h!]
\centering
\begin{subfigure}[t]{0.48\textwidth}
\centering
\includegraphics[width=\textwidth]{figures/trefoil_link_1_1_1_2.png}
\caption*{$A = \begin{pmatrix}1 & 1 \\ 1 & 2\end{pmatrix}$}
\end{subfigure}
\hfill
\begin{subfigure}[t]{0.48\textwidth}
\centering
\includegraphics[width=\textwidth]{figures/trefoil_link_2_3_5_8.png}
\caption*{$A = \begin{pmatrix}2 & 3 \\ 5 & 8\end{pmatrix}$}
\end{subfigure}
\begin{subfigure}[t]{0.48\textwidth}
\centering
\includegraphics[width=\textwidth]{figures/trefoil_link_13_8_21_13.png}
\caption*{$A = \begin{pmatrix}13 & 8 \\ 21 & 13\end{pmatrix}$}
\end{subfigure}
\hfill
\begin{subfigure}[t]{0.48\textwidth}
\centering
\includegraphics[width=\textwidth]{figures/trefoil_link_43_163_67_254.png}
\caption*{$A = \begin{pmatrix}43 & 163\\ 67 & 254\end{pmatrix}$}
\end{subfigure}
\caption{A matrix $A \in \PSLZ$ determines a periodic orbit of a flow in $\LS$, which in turn determines a modular trefoil link $k_A$ in $S^3$. Here, we visualize $k_A$ for several values of $A$.}
\label{fig:trefoil_links}
\end{figure}
The periodic orbits of the modular flow provide a rich variety of trefoil links.
One classically studied invariant of two-component links is their \defnphrase{linking number}, an integer associated with a particular link. The linking number of a two-component link can be defined computationally as follows:
\begin{enumerate}
\item Color the components of the link red and blue, respectively, and give each component an orientation.
\item Project the colored, oriented link onto the plane. Any regular projection will do.
\item Assign each crossing of a red strand over blue strand a number, either $1$ or $-1$ according to the diagram in Figure~\ref{fig:linking_number}.
\item Sum the integers at all crossings to obtain the linking number.
\end{enumerate}
\begin{figure}[t]
\centering
\includegraphics[width=0.3\linewidth]{figures/linking_number.pdf}
\caption{Diagrammatic rules for computing the linking number of link.}
\label{fig:linking_number}
\end{figure}
\begin{figure}[b]
\centering
\includegraphics[width=0.3\linewidth]{figures/linking_number_example.pdf}
\caption{A simple link. The red component crosses over the blue component exactly once. According to Figure~\ref{fig:linking_number}, this link has linking number 1.}
\label{fig:linking_number_example}
\end{figure}
A simple example is shown in Figure~\ref{fig:linking_number_example}.
For a more complex example, consider the first link in Figure~\ref{fig:trefoil_links}; it has linking number 0.
In \cite{ghys2007}, Ghys proves a fascinating result relating the linking numbers of modular trefoil links to a famous arithmetic function.
To understand this function and subsequent result, we first need to understand a few facts about $\SLZ$ and its relative, $\PSLZ$.
First, we will look at the basic building block of each group---their generators.
\begin{prop}
The group $\SLZ$ is generated by the matrices
\begin{equation*}
S = \begin{pmatrix}
0 & -1 \\
1 & 0
\end{pmatrix},\
T = \begin{pmatrix}
1 & 1 \\
0 & 1
\end{pmatrix}
\end{equation*}
\end{prop}
\begin{proof}
Let $\alpha \in \SLZ$.
First we show that there exists some $\gamma \in \ang{S,T}$ such that $\gamma\alpha$ is a matrix with lower left entry 0.
Note that for all $n \in \Z$,
\begin{equation*}
\begin{pmatrix}
1 & n \\
0 & 1
\end{pmatrix}
= \begin{pmatrix}
1 & 1 \\
0 & 1
\end{pmatrix}^n
= T^n \in \ang{S,T}
\end{equation*}
Then for any matrix $\begin{psmallmatrix}a & b \\ c & d\end{psmallmatrix}$, we have the identities
\begin{equation*}
S \begin{pmatrix}
a & b \\
c & d
\end{pmatrix}
= \begin{pmatrix}
-c & -d \\
a & b
\end{pmatrix},\ \
T^n \begin{pmatrix}
a & b \\
c & d
\end{pmatrix}
= \begin{pmatrix}
a + nc & b + nd \\
c & d
\end{pmatrix}
\end{equation*}
Now, let $\alpha = \begin{psmallmatrix}a & b \\ c & d\end{psmallmatrix}$.
If $c = 0$, then $\gamma = I$ and we are done.
So suppose $c \neq 0$.
If $|a| < |c|$, left multiply $\alpha$ by $S$.
Then $|a| \geq |c|$, so we can divide $a$ by $c$.
Then we have $a = qc + r$ with $0 \leq r < |c|$.
Then $T^{-q} \alpha$ has upper left entry $r = a - qc$, which has absolute value smaller then the lower left entry of $T^{-q} \alpha$.
Applying $S$ to $T^{-q} \alpha$ switches the upper and lower left entries, and we can apply the division algorithm again.
After each iteration of this process, we obtain a matrix with lower left entry whose absolute value is strictly less than before.
So eventually we have a word $\gamma$ in $S$ and $T$ such that $\gamma\alpha$ has lower left entry 0.
To summarize, at the end of the following algorithm we have a word $\gamma$ in $S$ and $T$ such that $\gamma \alpha$ has lower-left entry 0.
\vspace{1em}
\begin{algorithmic}
\State $\gamma \gets I$
\If{$(\gamma \alpha)_{0,0} > (\gamma \alpha)_{1,0}$}
$\gamma \gets S \gamma$
\EndIf
\While{$(\gamma \alpha)_{1,0} \neq 0$}
\State $q \gets a \ // \ c$
\State $r \gets a\ \% \ c$
\State $\gamma \gets T^{-q} \gamma$
\If{$(\gamma \alpha)_{0,0} > (\gamma \alpha)_{1,0}$}
$\gamma \gets S \gamma$
\EndIf
\EndWhile
\end{algorithmic}
Since the determinant is multiplicative, $\gamma \alpha$ also has determinant 1.
Hence $\gamma\alpha$ must be of the form $\begin{psmallmatrix}1 & n \\ 0 & 1\end{psmallmatrix}$ or $\begin{psmallmatrix}-1 & n \\ 0 & -1\end{psmallmatrix}$ for some $n \in \Z$.
So $\gamma \alpha = \pm T^n$, which means that $\alpha = \pm \gamma^{-1} T^n$.
Both $\gamma^{-1} T^n$ and $-\gamma^{-1} T^n = -I \gamma^{-1} T^n = S^2 \gamma^{-1} T^n$ are in $\ang{S,T}$.
Thus we have taken an arbitrary $\alpha \in \SLZ$ and represented it as an element in $\ang{S,T}$, thus $\SLZ \subseteq \ang{S,T}$.
The reverse inclusion is apparent.
\end{proof}
Given the generators of a group, we can of course write down every element in the group as a word in those generators. We now describe more precisely the structure of elements of $\PSLZ$ as generator words.
\begin{lemma}\label{lemma:pslgenerators}
The group $\PSLZ = \ang{S, T} = \ang{U,V} = \ang{P,Q}$, where
\begin{equation*}
U = \begin{pmatrix}
0 & 1 \\
-1 & 0
\end{pmatrix},\
V = \begin{pmatrix}
1 & -1 \\
1 & 0
\end{pmatrix},\
P = \begin{pmatrix}
1 & 1 \\
0 & 1
\end{pmatrix},\
Q = \begin{pmatrix}
1 & 0 \\
1 & 1
\end{pmatrix}
\end{equation*}
and $S$ and $T$ are as before.
\end{lemma}
\begin{proof}
The group $\PSLZ$ is all integer matrices $A$ such that $\det A = 1$, subject to the identification $A = -A$.
Since $\SLZ = \ang{S,T}$ and any such $A$ is in $\SLZ$, we have that $S$ and $T$ generate $\PSLZ$ as well.
In $\PSLZ$, we also have that $S = U = P^{-1}QP^{-1}$ and $T = VU = P$.
Thus $\PSLZ = \ang{S,T} = \ang{U,V} = \ang{P,Q}$.
\end{proof}
\begin{lemma}
Any element of $\PSLZ$ can be written in shortest form as
\begin{equation}\label{eq:pslreducedform}
V^{\varepsilon_0} U V^{\varepsilon_1} U \cdots V^{\varepsilon_{n}} U V^{\varepsilon_{n+1}}
\end{equation}
where $\varepsilon_0, \varepsilon_{n+1} \in \{-1,0,1\}$ and $\varepsilon_i \in \{-1,1\}$ for $1 \leq i \leq n$.
\end{lemma}
\begin{proof}
Let $A \in \PSLZ$.
Since $\PSLZ = \ang{U,V}$, we have that $A$ can be written in shortest form as
\begin{equation*}
V^{\varepsilon_0} U^{\varepsilon_1} V^{\varepsilon_2} \cdots U^{\varepsilon_{n}} V^{\varepsilon_{n+1}}
\end{equation*}
where $\varepsilon_i \neq 0$ for all $ 1 \leq i \leq n$.
But we also have that
\begin{equation*}
U^2 = V^3 = \begin{pmatrix}
-1 & 0 \\
0 & -1
\end{pmatrix}
= I.
\end{equation*}
Thus $\varepsilon_i \in \Z / 3 \Z$ for all $i$, and $\varepsilon_i \not\equiv 0 \pmod 3$ for $1 \leq i \leq n$.
And since $V^2 = V^{-1}$, we have that
\begin{equation*}
A = V^{\varepsilon_0} U V^{\varepsilon_1} U \cdots V^{\varepsilon_n} U V^{\varepsilon_{n+1}}
\end{equation*}
where $\varepsilon_i = \pm 1$ for $1 \leq i \leq n$, the desired result.
\end{proof}
The representation in \eqref{eq:pslreducedform} ends up being unique; for details see \cite[12]{rankin1977modular}.
Now, recall that the hyperbolic elements of $\PSLZ$ are those with trace greater than 2, and that we are interested in the hyperbolic elements because the conjugacy classes of hyperbolic elements are in bijection with the periodic orbits of the modular flow.
A result of Rademacher can help us say more about the generator words of hyperbolic elements in particular.
\begin{lemma}\label{lemma:tr_is_conj_invariant}
Let $A, B \in \PSLZ$.
If $\tr A \neq \tr B$, then $A$ is not conjugate to $B$.
\end{lemma}
\begin{proof}
Let
\begin{equation*}
A = \begin{pmatrix}
a_1 & a_2 \\
a_3 & a_4
\end{pmatrix}, \quad B = \begin{pmatrix}
b_1 & b_2 \\
b_3 & b_4
\end{pmatrix}, \quad \text{and} \quad C = \begin{pmatrix}
c_1 & c_2 \\
c_3 & c_4
\end{pmatrix}
\end{equation*}
be elements of $\PSLZ$ such that $\tr A \neq \tr B$, and assume for the sake of contradiction that $A$ is conjugate to $B$ by $C$.
Then
\begin{equation*}
\begin{array}{l r c l}
& B &=& CAC^{-1} \\[0.2em]
\implies & \begin{pmatrix}
b_1 & b_2 \\
b_3 & b_4
\end{pmatrix} &=&
\begin{pmatrix}
c_1 & c_2 \\
c_3 & c_4
\end{pmatrix}
\begin{pmatrix}
a_1 & a_2 \\
a_3 & a_4
\end{pmatrix}
\begin{pmatrix}
c_4 & -c_2 \\
-c_3 & c_1
\end{pmatrix} \\[1em]
\implies & b_1 + b_4 &=& (a_1 + a_4)(c_1 c_4 - c_2 c_3) \\[0.2em]
\implies & b_1 + b_4 &=& (a_1 + a_4)\det(C) \\[0.2em]
\implies & b_1 + b_4 &=& (a_1 + a_4)\\[0.2em]
\implies & \tr B &=& \tr A.
\end{array}
\end{equation*}
Thus we have a contradiction, and so $A$ is not conjugate to $B$.
\end{proof}
\begin{prop}[Rademacher]
Let $A \in \PSLZ$.
The conjugacy class of $A$ has a shortest representative either of the form
\begin{equation}\label{eq:uvform_elliptic}
\text{$U$, $V$, or $V^{-1}$}
\end{equation}
or of the form
\begin{equation}\label{eq:uvform}
U V^{\varepsilon_1} U V^{\varepsilon_2} \cdots V^{\varepsilon_{n}}
\end{equation}
where $\varepsilon_i \in \{-1,\, 1\}$ for $1 \leq i \leq n$.
Any cyclic permutation of the factors in such a representation remains in the same conjugacy class, thus these representatives are considered only up to cyclic permutation.
We call the form \eqref{eq:uvform} a \defnphrase{$UV$ decomposition of the conjugacy class of $A$}.
\end{prop}
\begin{proof}