-
Notifications
You must be signed in to change notification settings - Fork 10
/
ph229-notes.tex
executable file
·6499 lines (5635 loc) · 436 KB
/
ph229-notes.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{ws-rv9x6}
\usepackage{ws-rv-thm}
\usepackage[square]{ws-rv-van}
\usepackage{simplewick}
\usepackage{relsize}
\usepackage{alphalph}
\usepackage[vcentermath]{youngtab}
\newcommand{\myng}[1]{\,{\tiny\yng #1}\,}
\newif\ifarxivsubmission
\arxivsubmissiontrue
\ifarxivsubmission
\usepackage{color}
\definecolor{darkblue}{rgb}{0.1,0.1,.7}
\usepackage[colorlinks, linkcolor=darkblue, citecolor=darkblue, urlcolor=darkblue, linktocpage]{hyperref}
\crop[off]{}
\makeatletter
\def\@makechapterhead#1{%
{\vbox to 110pt{% %rvs
% \refstepcounter{chapter}%
\def\thefootnote{\@fnsymbol\c@footnote}%
% \addcontentsline{toc}{chapter}{\outchapter#1}
\vspace*{37pt} %VSPACE FROM TRIM SIZE
\parindent\z@\raggedright\reset@font
{\centering{%{\CNfont Chapter~\thechapter\par}%
\vskip 0.25in
\vbox{
\CTfont #1\par
}\par}\par}\nobreak\vfill}}}% %rvs
\makeatother
\else
\usepackage[]{hyperref}
\fi
\makeindex
%\newcommand\be{\begin{eqnarray}}
%\newcommand\ee{\end{eqnarray}}
\def\be#1\ee{\begin{align}#1\end{align}}
\newcommand\f\phi
\newcommand\cO{\mathcal{O}}
\newcommand\p[1]{\left(#1\right)}
\newcommand\ptl\partial
\newcommand\e\epsilon
\newcommand\<\langle
\renewcommand\>\rangle
\newcommand\Z{\mathbb{Z}}
\newcommand\de\delta
\newcommand\R{\mathbb{R}}
\newcommand\bx{\mathbf{x}}
\newcommand\nn{\nonumber}
\renewcommand\.{\cdot}
\newcommand\x\times
\newcommand\pdr[2]{\frac{\partial #1}{\partial #2}}
\newcommand\s\sigma
\newcommand\SO{\mathrm{SO}}
\newcommand\De{\Delta}
\newcommand\cS{\mathcal{S}}
\newcommand\oo\infty
\newcommand\SU{\mathrm{SU}}
\newcommand\cH{\mathcal{H}}
\newcommand\bn{\mathbf{n}}
\newcommand\bP{\mathbf{P}}
\renewcommand\b\beta
\renewcommand\a\alpha
\newcommand\Tr{\mathrm{Tr}}
\renewcommand\l\lambda
\newcommand\cL{\mathcal{L}}
\newcommand\cD{\mathcal{D}}
\renewcommand\th{\theta}
\newcommand\tl[1]{\widetilde{#1}}
\newcommand\GeV{\mathrm{GeV}}
\newcommand\g\gamma
\newcommand\arXiv[1]{{\tt arXiv:#1}}
\newcommand\z\zeta
\renewcommand\hat[1]{\widehat{#1}}
\newcommand\bs{\mathbf{s}}
\newcommand\draftnote[1]{{\color{blue} #1}}
\newcommand\C{\mathbb{C}}
\renewcommand\bar[1]{\overline{#1}}
\newcommand\SL{\mathrm{SL}}
\newcommand\cM{\mathcal{M}}
\newcommand\cN{\mathcal{N}}
\newcommand\cA{\mathcal{A}}
\newcommand\cF{\mathcal{F}}
\newcommand\w{\omega}
\newcommand\stoplecture{
\stepcounter{LectureCounter}
\hfill \draftnote{\it End of lecture \theLectureCounter}
}
\newcounter{LectureCounter}
\newtheorem{exercise}{Exercise}[section]
\let\Exercisefont\itshape
\def\Exerciseheadfont{\bfseries}
\numberwithin{equation}{section}
\begin{document}
\ifarxivsubmission
\chapter{Phys 229ab Advanced Mathematical Methods:\ \ \\
Conformal Field Theory
}\label{ch:bootstrap}
\else
\chapter{The Conformal Bootstrap}
\fi
\author[David Simmons-Duffin]{David Simmons-Duffin}
\address{Caltech,
\begin{abstract}
Notes for Physics 229, 2017-2018. These notes are in progress. They were last updated on \today. In a few places, there are reminders \draftnote{in blue} to the author to add additional discussion/comments. There may also be empty citations and missing figures. Some of the discussion is not original and is borrowed from various sources. I have tried to indicate with a footnote when a section closely follows a particular reference.
\end{abstract}
\body
\makeatletter
\newalphalph{\alphmult}[mult]{\@alph}{26}
\makeatother
\renewcommand{\thefootnote}{\arabic{footnote}}
\ifarxivsubmission
\thispagestyle{empty}
\makeatletter
\renewcommand*\l@chapter[2]{}
\renewcommand*\l@schapter[2]{}
\renewcommand*\l@author[2]{}
\makeatother
\newpage
\setcounter{page}{1}
\setcounter{tocdepth}{3}
\smalltoc
\tableofcontents
\fi
\ifarxivsubmission
\newpage
\fi
\section*{Resources}
The introductory part of these notes is heavily based on Silviu Pufu's Bootstrap 2017 lectures \cite{Pufu2017}, John Cardy's book {\it Scaling and Renormalization in Statistical Physics\/} \cite{Cardy:1996xt}, Slava Rychkov's EPFL lectures on CFT \cite{Rychkov:2016iqz}, and John McGreevy's lectures on QFT \cite{McGreevy2017}. The core material on axiomatic CFT is based on my TASI lectures \cite{}. The section on 2d CFTs is based on material in Polchinski's String Theory book \cite{} and di Francesco et.\ al.\ \cite{}.
\section{Introduction}
\subsection{QFT and emergent symmetry}
Quantum Field Theory is a universal language for theoretical physics. It shows up in many different settings, for example
\begin{itemize}
\item statistical physics,
\item condensed matter physics,
\item particle physics (SM and beyond),
\item string theory/holography.
\end{itemize}
The microscopic physics in these settings can be quite complicated. However, the macroscopic physics sometimes displays extra ``emergent" symmetries that can help us do computations.
For example, in condensed matter physics, we are interested in describing a material made up of atoms with some lattice spacing $a$.\footnote{We ignore the possibility of lattice defects for the moment.} We refer to the detailed lattice system as the ``microscopic theory." Quantum field theory is a good description at distances much larger than the lattice spacing $x\gg a$, or equivalently energy/momenta much lower than the UV cutoff $\Lambda_{UV}=1/a$.\footnote{QFTs that are valid below a UV cutoff are often called ``effective field theories" (EFTs).} At these large scales, the discrete translation symmetry of the lattice becomes a continuous symmetry.
Similarly, in statistical physics, particle physics, and string theory, QFT can describe distance scales much larger than the characteristic scales of the microscopic theory.
Statistical systems are described by QFTs in Euclidean signature, e.g.\ on $\R^d$. Such QFTs capture properties of the equilibrium state. By contrast, condensed matter and particle systems are described by QFTs in Lorentzian signature, e.g.\ on $\R^{d-1,1}$. Such QFTs encode time-dependent quantum dynamics.
We will be interested in QFTs with rotational symmetry, by which we mean $\SO(d)$ symmetry in Euclidean signature and $\SO(d-1,1)$ symmetry in Lorentzian signature. In particle physics and string theory, this symmetry is built into the microscopic theory. However in lattice systems, rotational symmetry is emergent. This means that correlation functions become rotationally invariant in the limit of large distances, even though microscopic correlation functions are not rotationally-invariant.\footnote{Emergent rotational symmetry is very familiar: weÊ often cannot determine the orientation of a microscopic lattice using macroscopic observations. Some examples of materials {\it without\/} emergent $\SO(d)$ symmetry are crystals like salt. A very exotic example is the Haah code \cite{}.} In particular, for condensed matter systems, the effective ``speed of light" associated with $\SO(d-1,1)$-invariance is an emergent property and has nothing to do with the speed of actual light. (We will see some explicit examples later.)
Under general conditions, $\SO(d)$-invariant Euclidean QFTs are in one-to-one correspondence with $\SO(d-1,1)$-invariant Lorentzian QFTs. The map between them is called Wick rotation, and we will discuss it in detail. Because of this correspondence, we can focus mostly on Euclidean QFTs, and later understand Lorentzian QFTs by Wick rotating what we learned in Euclidean signature.
\subsection{The mass gap and critical points}
Thus, we are interested in theories with Poincare symmetry
\be
G_\mathrm{Poincare} &= \R^d \rtimes \SO(d).
\ee
From the point of view of long-distance physics, the most important property of a Poincare-invariant theory is its mass gap
\be
m_\mathrm{gap} &= E_1-E_0 \geq 0,
\ee
where $E_0,E_1$ are the energies of the ground state and first excited state, respectively.\footnote{Some QFTs with topological order can have multiple degenerate ground states.} Theories with nonzero $m_\mathrm{gap}$ are called ``gapped."
To understand why $m_\mathrm{gap}$ is important, let us study a two-point function of a scalar local operator $\f(x)=\f(x^0,\mathbf{x})$. We can ensure that $\f(0)$ has vanishing vacuum expectation value by subtracting off an appropriate multiple of the unit operator. By Poincare invariance, it suffices to consider $\<0|\f(x^0,\mathbf{0}) \f(0)|0\>$ with $x^0>0$. The two-point function is then given by
\be
\label{eq:completeset}
\<0|\f(x^0,\mathbf{0})\f(0)|0\> &= \<0|\f(0) e^{-H x^0} \f(0)|0\> \nn\\
&= \sum_\psi |\<0|\f(0)|\psi\>|^2 e^{-E_\psi x^0}.
\ee
Here, $H$ is the Hamiltonian with the vacuum energy subtracted off, and $\psi$ runs over an orthonormal basis of eigenstates of $H$. To get the right-hand side, we have used $\f(x^0,\mathbf{0}) = e^{H x^0} \f(0,\mathbf{0}) e^{-H x^0}$ and $H|0\>=0$.\footnote{Note that the Euclidean time-evolution operator is $e^{-H x^0}$ as opposed to the familiar $e^{-iHt}$ in Lorentzian signature. They are related by Wick rotation $x^0 = i t$. We will discuss this in much more detail in later sections.}
The key point is that the operator $e^{-H x^0}$ exponentially damps states with energy $E_\psi\gg 1/x^0$.
%%We have the Kallen-Lehmann spectral representation
%%\be
%%\<\f(x)\f(y)\> &= \int dm^2 \rho(m^2) G_F(x-y,m),
%%\ee
%%where $G_F(x,m^2)$ is the Feynman propagator with mass $m$. Here, $\rho(m^2)\geq 0$ is an effective density of states with mass $m$. If we demand that $\f(0)$ have vanishing vacuum expectation value (by subtracting off an appropriate multiple of the unit operator), then the support of $\rho(m^2)$ is contained in $[m_\mathrm{gap}^2,\oo)$.
%%
%% For simplicity, consider $d=3$ where the Feynman propagator simplifies
%%\be
%%G_F(x,m) &= \frac{1}{4\pi |x|} e^{-m|x|} \qquad (d=3).
%%\ee
%The key point is that $G_F(x,m)$ falls off exponentially as $e^{-m|x|}$ in Euclidean signature. At large $|x|$, the two-point function is then dominated by contributions with the smallest $m$, and thus falls off as $e^{-m_\mathrm{gap} |x|}/|x|$ if $m_\mathrm{gap}$ is nonzero.\footnote{It could fall faster if the operator $\f(x)$ has zero amplitude to create the lightest state in the theory.}
%
%A more general way to see the exponential falloff is to write the two-point function as
%\be
%\<0|\f(x^0,\mathbf{0}) \f(0)|0\> &= \<0|\f(0)e^{-H x^0} \f(0)|0\>,
%\ee
%where $H$ is the Hamiltonian (with the ground state energy subtracted off), and we have taken the points to be separated only in the $x^0$ direction for simplicity.\footnote{Note that in Euclidean theories, the time-evolution operator is $e^{-H x^0}$ as opposed to $e^{-iHt}$ familiar from quantum mechanics. We will discuss this in much more detail later in the course.}
At large $x^0$, the correlator is dominated by $\psi$ with the smallest nonzero eigenvalue of $H$, which is $m_\mathrm{gap}$.\footnote{Note that the vacuum does not contribute as an intermediate state because we have demanded $\<0|\f(0)|0\>=0$.}
Thus, when $m_\mathrm{gap}$ is nonzero, correlation functions of local operators fall off at least as fast as $e^{-|x|/\xi}$, where $\xi\equiv 1/m_\mathrm{gap}$ is called the ``correlation length." Generic statistical and condensed matter systems have microscopic correlation lengths $\xi \sim a$, or equivalently $m_\mathrm{gap} \sim \Lambda_{UV}$. At long distances, they are described by QFTs whose local correlation functions vanish, called topological quantum field theories (TQFTs).
%Generic statistical and condensed-matter systems are gapped. In such systems, correlations at scales $x\gg \l$ are effectively zero.\footnote{Gapped theories are described by topological quantum field theories (TQFTs) and are interesting in their own right. However, we will not focus on TQFTs in this course.}
However, sometimes by tuning parameters in the microscopic Hamiltonian, we can make $m_\mathrm{gap}$ much smaller than $\Lambda_{UV}$, and even arrange for $m_\mathrm{gap}$ to vanish. Points in parameter space where $m_\mathrm{gap}=0$ are called critical points.
At a critical point, the system experiences a phase transition, and develops nonzero correlations at arbitrarily long distances.\footnote{If the Standard Model were like a generic condensed matter system, we might expect $m_\mathrm{gap}$ to be close to the UV cutoff, which is perhaps the GUT scale $10^{15}\GeV$ or Planck scale $10^{18}\GeV$. The hierarchy problem is the problem of explaining why the Standard Model is so close to a critical point.}
Long-distance correlation functions at critical points have no intrinsic length-scale because all memory of dimensionful microscopic quantities (like the lattice spacing $a$) disappears when distances become arbitrarily large. For example, critical two-point functions behave as pure power laws
\be
\label{eq:criticaltwopt}
\<\f(x)\f(0)\> &= \frac{C}{|x|^{2\De}} \qquad(\textrm{critical point, }x\gg a),
\ee
where $C$ and $\De$ are constants depending on $\phi$. The quantity $\De$ is called the scaling dimension of $\f$.
\subsection{Scaling and conformal symmetry}
A more precise way to state the lack of an intrinsic length scale is to say that theories with $m_\mathrm{gap}=0$ have an emergent symmetry under rescaling
\be
\label{eq:rescalingsymmetry}
x^\mu &\to \l x^\mu\qquad (\l>0).
\ee
Rescaling also acts on the operators\footnote{We will give a precise definition of what this means in section~\ref{}.}
\be
\f(x) &\to \l^\De \f(\l x),
\ee
and it is easy to check that (\ref{eq:criticaltwopt}) is invariant under this transformation.
Under very general conditions (that we will discuss), critical points also display less obvious emergent symmetries called conformal transformations. A conformal transformation $x\to x'(x)$ is a map that looks like a rotation and rescaling near each point,
\be
\pdr{x'^\mu}{x^\nu} &= \Omega(x) R^{\mu}{}_\nu(x),\qquad R^{\mu}{}_\nu\in \SO(d).
\ee
An example is a special conformal transformation,
%a less-obvious emergent symmetry under special conformal transformations
\be
\label{eq:specialconformalsymmetry}
x^\mu \to \frac{x^\mu - b^\mu x^2}{1-2 b\.x + b^2 x^2}\qquad(b\in \R^d).
\ee
In 2 dimensions, there are more exotic examples, like the one pictured in figure~\ref{}.
Here is some rough intuition for why critical points display conformal symmetry. We know that a critical theory is invariant under rescalings and rotations. If the theory is also local, in the sense that degrees of freedom at a point only interact directly with other degrees of freedom at nearby points, then the theory should also be invariant under transformations that locally look like a rescaling and rotation. This is the defining property of a conformal transformation. Turning this rough intuition into a theorem is a difficult problem (that we will discuss in more detail later). However, it seems to be true in a wide class of systems.
QFTs that are invariant under conformal symmetry are called conformal field theories (CFTs). To summarize, CFTs describe critical points where $m_\mathrm{gap}=0$. One can also understand the neighborhood of a critical point where $m_\mathrm{gap}$ is nonzero (but still $m_\mathrm{gap} \ll \Lambda_{UV}$) by studying perturbations of the associated CFT.
\subsection[Examples of critical points]{Examples of critical points\footnote{Sources for this section: Pufu \cite{Pufu2017}, Cardy \cite{Cardy:1996xt}.}}
So far our discussion has been very abstract, so let us introduce some examples.
%Through examples, we will try to explain
%\begin{itemize}
%\item Why statistical systems are described by Euclidean QFT,
%\item How different microscopic theories are described by the same CFT (universality),
%\item How to tune $m_\mathrm{gap}\to 0$ in a solvable model,
%\item How to study (some) CFTs using conventional QFT tools like perturbation theory.
%\end{itemize}
One of our goals will be to infer from examples a set of axioms that CFTs should satisfy. We will then study these axioms in the next part of the course. Another goal will be to be more precise about how and why statistical and condensed matter systems are described by quantum field theory, and how critical points come about.
\subsubsection{Magnets}
Our first examples of critical points occur in magnets. Given a magnet with temperature $T$, we can apply a magnetic field $\vec{H}$ and measure the magnetization $\vec{M}$. There are three main types of magnets in 3-dimensions, which are distinguished by their symmetries:\footnote{Here, we mean non-spacetime symmetries, usually called ``global" or ``flavor" symmetries. The emergent spacetime symmetry group is still the Poincare group, or the conformal group at the critical point.}
\begin{itemize}
\item {\it Uni-axial magnet}: individual magnetic moments $\vec\mu$ are confined to lie along a fixed axis. Uni-axial magnets have an $O(1)=\Z_2$ symmetry under which $\vec H \to -\vec H$ and $\vec M \to -\vec M$.
\item {\it XY magnet}: magnetic moments $\vec\mu$ are oriented in a plane. Such magnets have an $O(2)$ symmetry under which $\vec H$ and $\vec M$ rotate in the plane.
\item {\it Heisenberg magnet}: magnetic moments $\vec \mu$ are unconstrained. Such magnets have $O(3)$ symmetry under which $\vec H$ and $\vec M$ transform in the vector representation.
\end{itemize}
For the moment, we will focus on the simplest case of uni-axial magnets. We denote the projections of $\vec H,\vec M$ onto the appropriate axis by $H,M$.
In experiments, we observe the following. There exists a ``critical temperature" $T_c$, such that
\begin{itemize}
\item For $T<T_c$, the preferred state of the magnet has nonzero magnetization $M\neq 0$ when $H=0$. In other words, the $\Z_2$ symmetry is spontaneously broken.
\item For $T>T_c$, the magnet has $M=0$ when $H=0$, i.e.\ the $\Z_2$ symmetry is unbroken.
\end{itemize}
The corresponding phase diagram is pictured in figure~\ref{}. The point
\be
H = 0,\ T=T_c
\ee
is a critical point, and is described by a CFT at long distances. To reach the critical point, we must tune two parameters: $H$ and $T$. Tuning $H=0$ is easy because that is where the microscopic theory has $\Z_2$ symmetry. However, the value of $T_c$ depends on the specific material.
In more detail, the behavior of the magnetization in different phases is shown in figure~\ref{}. Close to $T_c$, observables exhibit so-called ``scaling" behavior, characterized by various critical exponents. Let us define the dimensionless couplings
\be
\label{eq:dimensionlesscouplings}
t \equiv \frac{T-T_c}{T_c},\qquad h=\frac{H}{k_B T}.
\ee
Some examples of critical exponents are
\begin{itemize}
\item $\a$: the heat capacity at $h=0$ behaves as
\be
C &= \frac{\ptl^2 F}{\ptl T^2} \propto |t|^{-\a}.
\ee
(Here $F$ is the free-energy.)
\item $\b$: the spontaneous magnetization behaves as
\be
\lim_{H\to 0^+} M &\propto (-t)^\b.
\ee
\item $\gamma$: the zero-field susceptibility behaves as
\be
\chi &= \left.\pdr{M}{H}\right|_{H=0} \propto |t|^{-\g}.
\ee
\item $\de$: the magnetization at $T=T_c$ behaves as
\be
|M|&\propto|h|^{1/\de}.
\ee
\item $\nu$ and $\eta$: the correlation length $\xi$ can be measured by studying a two-point correlation functions of spins
\be
G(x) &= \<s(x)s(0)\> - \<s(0)\>^2.
\ee
Away from the critical point, $G(x)\sim e^{-|x|/\xi}$ decays exponentially. However, as $t\to 0$, the correlation length diverges as
\be
\label{eq:correlationlengthexponent}
\xi &\propto |t|^{-\nu}.
\ee
Equivalently, the mass-gap goes to zero as $m_\mathrm{gap}\propto |t|^\nu$.
Precisely at $t=0$, the two-point function takes the form
\be
G(x) &\propto \frac{1}{|x|^{d-2+\eta}},
\ee
i.e.\ the spin operator has dimension $\Delta_s = \frac{d-2}{2}+\frac \eta 2$.
\end{itemize}
Don't worry, I can't keep track of all these critical exponents either. We will see shortly that all of this behavior can be explained using effective field theory, scaling symmetry, and dimensional analysis.
Now, here is an amazing fact:
\begin{claim}
We find the same critical exponents in many different uni-axial magnets, regardless of what material they're made of.
\end{claim}
In fact, critical uni-axial magnets are all described by the same scale-invariant QFT at long distances. This phenomenon is called ``critical universality."
\subsubsection{Liquid-vapor transitions}
Other critical points appear in liquid-vapor transitions. For example, the phase-diagram of water is pictured in figure~\ref{}. Near room temperatures and pressures, there is a sharp distinction between the liquid and gas phases. However, at higher temperatures and pressures, the distinction between liquid and gas disappears at a critical point $(T_c,P_c)$. For example, in water $T_c=647\,\mathrm{K}$, $P_c=374\,\mathrm{Atm}$.
Note that the critical points of magnets and water are both obtained by tuning two parameters. Comparing neighborhoods of the critical points in figures~\ref{} and \ref{}, we can make the following rough analogy between water and magnets:
\be
P-P_c &\sim H, \nn\\
\rho-\rho_c &\sim M.
\ee
where $\rho$ is the density and $\rho_c$ is the critical density.\footnote{Water does not have a microscopic $\Z_2$ symmetry, but it turns out that one emerges near the critical point. Roughly speaking, the $\Z_2$ switches the liquid (high-density) and gas (low-density) phases. To make a more precise analogy, we should identify $M$ with the combination of $\rho$ and $P$ that flips sign under the emergent $\Z_2$.}
In measurements of critical water, we again observe scaling behavior. For example, the heat capacity behaves as
\be
C &\sim |t|^{-\a},
\ee
where $t$ is again given by (\ref{eq:dimensionlesscouplings}).
Additionally, the difference in density between the liquid and gas phases behaves as
\be
\rho_\mathrm{liquid}-\rho_\mathrm{gas} &\sim (-t)^\b.
\ee
Amazingly,
\begin{claim}
Water and other liquid-vapor transitions have precisely the same critical exponents as uni-axial magnets.
\end{claim}
We say that liquid-vapor transitions are in the same ``universality class" as uni-axial magnets.
\stoplecture
\subsubsection{The Ising model}
The Ising model is a simplified model of a magnet that is still rich enough to capture its critical behavior. Its degrees of freedom are classical spins taking values $\pm 1$, with one spin for each site on a cubic lattice\footnote{The Ising model can also be formulated on other lattices, and in many cases these models lie in the same universality class.}
\be
s_i \in \{\pm 1\},\qquad i \in \Z^d.
\ee
The partition function is a sum over all configurations of spins, with Boltzmann weights that mimic the interactions between physical spins in a magnet,
\be
Z(K,h) &= \sum_{\{s_i\}} e^{-S[s]}, \nn\\
S[s] &= -K \sum_{\<ij\>} s_i s_j - h \sum_j s_j.
\ee
The notation $\<ij\>$ means that $i,j$ are neighboring lattice sites. (You can think of $K$ as the product $\b J$ where $\b$ is the inverse temperature and $J$ is the spin-spin interaction.) We can also compute correlation functions by inserting spins into the sum
\be
\<s_{i_1} \cdots s_{i_n}\> &= \frac{1}{Z}\sum_{\{s_i\}} s_{i_1} \cdots s_{i_n} e^{-S[s]}.
\ee
The ``magnetization" is proportional to the one-point function $\<s_i\>$ (which is independent of $i$).
When $K$ is positive, the term in $S[s]$ proportional to $K$ makes spins want to align. This term has a $\Z_2$ symmetry under which $s_i\to -s_i$ for all $i$. The term proportional to $h$ breaks this $\Z_2$ symmetry and causes spins to preferentially have the same sign as $h$. Both of these effects compete against statistical fluctuations.
In dimension $d \geq 2$, the Ising model famously exhibits a critical point at a special value $K=K_c$ and $h=0$. For $K>K_c$, spins spontaneously align and break the $\Z_2$ symmetry. For $K<K_c$, statistical fluctuations cause the spins to randomize and the $\Z_2$ is unbroken.
The Ising critical point displays precisely the same long-distance correlation functions and critical exponents as real uni-axial magnets and liquid-vapor systems. In fact, for our purposes, we can think of the Ising model as yet another physical system in the same universality class as these other systems. Its main distinguishing feature is that it is much simpler than e.g.\ actual water, and thus much easier to study.\footnote{There are many other abstract statistical lattice models in the same universality class as the Ising model, for example compass models \arXiv{1303.5922} and the Blume-Capel model \arXiv{1711.10946}. The Ising model on a cubic lattice is arguably the simplest, however the Blume-Capel model has some features that make it even better for simulation.} By studying the mathematics of the Ising model, we will be able to understand gaplessness and a few QFT/CFT axioms in a simple way. By the power of critical universality, quantities we compute in the abstract Ising model at $K=K_c$ agree exactly with interesting quantities in real physical systems.
\subsubsection{Continuum $\f^4$ theory}
The Ising lattice model is a good starting point for understanding why classical statistical systems are described by QFT at long distances. We can think of each configuration of spins as a map
\be
s: \Z^d \to \{\pm 1\},
\ee
and the partition function is a sum over all such maps. This is a discrete version of a path integral.
Typically, in QFT we integrate over continuous functions on a continuous space, e.g.\
\be
\f : \R^d \to \R.
\ee
If we study correlation functions of the Ising model at long distances (much larger than the lattice spacing), it's not hard to imagine that we can approximate the lattice as continuous $i\in\Z^d \to x\in\R^d$. Similarly, we can replace individual spins $s_i$ with average densities of spins $\f(x)$ in small neighborhoods, so that the effective spin at a point becomes a real number.
This suggests that the partition sum of the Ising model might be related to the path integral for a scalar field on $\R^d$, at least at long distances. That is, perhaps we can take
\be
\Z^d &\to \R^d\nn\\
s_i &\to \phi(x) \nn\\
\sum_{\{s_i\}} &\to \int D\phi \nn\\
e^{-S[s]} &\to e^{-S[\phi]},
\ee
for some continuum action $S[\f]$, without changing the long-distance behavior of correlation functions too much.
There is a non-rigorous procedure called ``taking the continuum limit" of a lattice model that tries to justify these replacements. However, it involves several uncontrolled approximations, and it is perhaps more honest to just study a continuum scalar theory and compare its behavior with the Ising model.
For concreteness, let us focus on 3-dimensions. The simplest interacting theory of a scalar field in 3d has Euclidean action
\be
\label{eq:continuumphifour}
S[\f] &= \int d^3 x \p{\frac 1 2 (\ptl \phi)^2 + \frac 1 2 m^2 \f^2 + g \f^4}.
\ee
Note that $\f$ has mass-dimension $1/2$, so that $m$ and $g$ both have mass-dimension $1$. Thus, $\f^2$ and $\f^4$ are relevant interactions, and we can think of this theory as an RG flow from the free-boson at high energies to something else at low energies.
Let us map out the phase diagram of this theory. Because $m$ and $g$ are the only dimensionful parameters, only their ratio matters.\footnote{To be more precise, there is also a UV cutoff $\Lambda$ that depends on how the theory is regularized. The quartic term gives rise to a linear divergence proportional to the $\f^2$ term, so physical masses are given by $m_\mathrm{phys}^2 = m^2 + \a g \Lambda$, with $\a$ a dimensionless number. (In some regularization schemes, like dimensional regularization, the linear divergence vanishes.) However, shifting $m^2$ is the only way the cutoff can enter a physical observable. Thus, our discussion becomes correct after the replacement $m^2 \to m^2_\mathrm{phys}$.} Suppose first that $m^2/g^2 \gg 0$ is large and positive. In this limit, the theory has a single massive vacuum with mass approximately $m_\mathrm{gap} \approx m$. Now suppose $m^2/g^2 \ll 0$ is very negative. In this case, the theory has two very massive vacuua that spontaneously break the $\Z_2$ symmetry, with $m_\mathrm{gap}\approx \sqrt {-2m^2}$.
If we start at very large $|m^2|$ (in either phase) and decrease $|m^2|$, $m_\mathrm{gap}$ decreases too. There exists a critical ratio $m^2/g^2 = r_c$ where $m_\mathrm{gap}\to 0$ and the theory is described by a nontrivial CFT. One can justify this claim using the $\e$-expansion and other perturbative techniques, as we will discuss, or from numerical simulations.
The phase diagram we have just described is the same as the Ising phase diagram. Indeed $\f^4$-theory is also in the Ising universality class.
Near the critical point, the $\f^4$ interaction cannot be treated as a small perturbation. More explicitly, because the coupling constant $g$ has mass-dimension 1, perturbation theory is really an expansion in $g x$, where $x$ is the characteristic length scale of the observable we are computing. In the long-distance limit $x\to \oo$ (where the CFT emerges), this expansion breaks down. Thus, the Ising CFT cannot be studied with traditional perturbation theory. We will later discuss two different perturbative (but uncontrolled) expansions for the Ising CFT: the $\e$-expansion and the large-$N$ expansion.
\begin{figure}
\begin{center}
\includegraphics[width=0.55\textwidth]{rgflows-ising.jpg}
\end{center}
\caption{Many microscopic theories can flow to the same IR CFT\@. We say that the theories are IR equivalent, or IR dual. \label{fig:rgflows}}
\end{figure}
The theory with action (\ref{eq:continuumphifour}) and $m^2/g^2=r_c$ is equivalent to the free boson at short distances (high energies/UV) and the Ising CFT at long distances (low energies/IR). The free boson is itself a CFT, so we have an RG flow from one CFT to another (figure \ref{fig:rgflows}). This construction, where we start with a UV CFT and perturb it so that it flows to an IR CFT, is one possible definition of a general QFT. In this definition, we must allow for the possibility that the IR theory is gapped, in which case $\mathrm{CFT}_\mathrm{IR}$ is a TQFT which is technically a special case of a CFT (where all local correlation functions are zero).
An RG flow from the free boson is perhaps the cleanest theoretical construction of the Ising CFT. It shows that the Ising CFT inherits properties of a continuum quantum field theory. For example, $\f^4$-theory has rotational invariance and $\Z_2$ symmetry, so the Ising CFT does too. The Ising lattice model is easier to simulate, but it does not have microscopic rotational symmetry. Meanwhile, water has rotational symmetry but no microscopic $\Z_2$ symmetry. We will learn a lot about the Ising CFT by going back and forth between different microscopic realizations.
\subsubsection{Other universality classes}
We have seen several different systems that fall into the Ising universality class. However, not every critical point is described by the Ising CFT. Another important class of 3d CFTs are the $O(N)$ models. These can be described as the critical point of a theory of $N$ bosons $\f_i$ ($i=1,\dots,N$) with a quartic interaction that respects an $O(N)$ global symmetry
\be
S &= \int d^3 x \p{\frac 1 2 \sum_i \ptl_\mu \f_i \ptl^\mu \f_i + \frac 1 2 m^2 \sum_i \f_i \f_i + g \p{\sum_i \f_i \f_i}^2}.
\ee
XY magnets are in the same universality class of the $O(2)$ model and Heisenberg magnets are in the same universality class of the $O(3)$ model.
\draftnote{Discuss non-uniqueness of UV Lagrangian description. An example is particle-vortex duality between the gauged $O(2)$ model and the $O(2)$ model. Distinguish between IR duality and exact duality.}
But this is just the tip of the iceberg --- there is a huge zoo of different types of critical points that we can construct in a variety of ways. One of our goals will be to better understand this zoo.
\section[Path integrals, quantization, and the quantum-classical correspondence]{Path integrals, quantization, and the quantum-classical correspondence\footnote{Sources for this section: McGreevy \cite{McGreevy2017}.}}
In this section, we explain in more detail why classical statistical systems are described by Euclidean quantum field theory, and also how they are related to quantum condensed-matter systems via Wick rotation. Along the way, we will introduce some important technical concepts in QFT, like cutting-and-gluing rules for path integrals and the idea of quantization.
\subsection{The 1d Ising model and the transfer matrix}
Let us start with the Ising lattice model in 1-dimension. For concreteness, we will study the theory on a periodic lattice with length $M$, so the spins $s_i$ are labeled by $i\in \Z_M$. The partition function is given by
\be
Z_1 &= \sum_{\{s_i=\pm 1\}} e^{-S[s]} \nn\\
S[s] &= -K \sum_{i=1}^{M} s_i s_{i+1} - h \sum_{i=1}^M s_i.
\ee
We refer to $S[s]$ as the ``action," even though it is equal to $\b H$, where $H$ is the classical Hamiltonian. This is because we would like to reserve the word Hamiltonian for a completely different object that will appear shortly.
As mentioned in the introduction, the partition sum should be thought of as a discrete version of a 1-dimensional path-integral. This 1-dimensional path integral can be computed by relating it to a 0-dimensional quantum theory. This is an example of the notion of {\it quantization}.
The key idea is to build up the partition sum by moving along the lattice site-by-site. Forget about periodicity for the moment, and consider the contribution to the partition function from spins $j<i$ for some fixed $i$,
\be
Z_\mathrm{partial}(i,s_i) &= \sum_{\{s_j:j<i\}} e^{K\sum_{j < i} s_j s_{j+1} + h \sum_{j< i} s_j}.
\ee
Because of the interaction term $s_{i-1}s_{i}$, we cannot do the sum over $\{s_1,\dots, s_{i-1}\}$ without specifying the spin $s_i$. Thus, we have a function of $s_i$. In short, $Z_\mathrm{partial}(i,s)$ is the partition function of the theory on the lattice $1\dots i$, with fixed boundary condition $s$ at site $i$.\footnote{We must also impose some boundary condition at site $1$. The precise choice is not important for this discussion, so we have left it implicit.}
Note that $Z_\mathrm{partial}(i+1,s_{i+1})$ can be related to $Z_\mathrm{partial}(i,s_{i})$ by inserting the remaining Boltzmann weights that depend on $s_i$ and performing the sum over $s_i=\pm 1$,
\be
\label{eq:onestepsum}
Z_\mathrm{partial}(i+1,s_{i+1}) &= \sum_{s_i = \pm 1} T(s_{i+1},s_i) Z_\mathrm{partial}(i,s_{i}),
\ee
where
\be
T(s_{i+1},s_i) &\equiv e^{K s_i s_{i+1} + h s_i}.
\ee
The key step is to recognize (\ref{eq:onestepsum}) as a discrete version of the Schrodinger equation in a 2-dimensional Hilbert space $\mathcal{H}$. This Hilbert space has basis $|s\>=|\pm 1\>$. The $T(s',s)$'s are elements of a $2\x 2$ matrix $\hat{T}$ acting on $\mathcal{H}$
%We can think of $T(s',s)$ as elements of a $2\x 2$ matrix and $Z_\mathrm{partial}(s)$ as components of a vector in a 2-dimensional Hilbert space,
\be
T(s',s) &= \<s'|\hat{T}|s\>,\qquad \hat{T} =
\begin{pmatrix}
e^{K+h} & e^{-K-h} \\
e^{-K+h} & e^{K-h}
\end{pmatrix},
\ee
and $Z_\mathrm{partial}(i,s)$ are the components of a vector $|\Psi_i\>\in \mathcal{H}$,
\be
Z_\mathrm{partial}(i,s) &= \<s|\Psi_i\>.
\ee
In this notation, (\ref{eq:onestepsum}) becomes
\be
\label{eq:discreteshrodinger}
|\Psi_{i+1}\> &= \hat{T} |\Psi_i\>.
\ee
The matrix $\hat{T}$ is called the ``transfer matrix", and it plays the role of a discrete time-translation operator. Here, $i$ should be thought of as a discrete Euclidean time coordinate.
To be explicit, the (integrated) Schrodinger equation in a quantum theory in Euclidean time is
\be
|\Psi(\tau+\Delta\tau)\> &= e^{-\De\tau \hat H} |\Psi(\tau)\>,
\ee
where $\tau$ is the Euclidean time coordinate, $\De\tau$ is some time-step, and $\hat H$ is the quantum Hamiltonian. Thus, the $1$-dimensional Ising lattice model is equivalent to a 2-state quantum theory with Hamiltonian
\be
\hat H &= -\frac{1}{\Delta\tau} \log \hat T.
\ee
When the lattice is periodic, the partition function is related to the transfer matrix by
\be
\label{eq:partitionfunctionastransfermatrix}
Z_1 %&= \sum_{\{s_i\}} T(s_M, s_{M-1}) T(s_{M-1},s_{M-2}) \cdots T(s_1,s_M) \nn\\
&= \sum_{\{s_i\}} \<s_M|\hat{T}|s_{M-1}\>\<s_{M-1}|\hat{T}|s_{M-2}\> \cdots \< s_1|\hat{T}|s_M\> \nn\\
&= \Tr(\hat{T}^M).
\ee
This is now easy to evaluate by diagonalizing $\hat T$,
\be
\Tr(\hat{T}^M) &= \l_+^M + \l_-^M,
\ee
where
\be
\l_\pm &= e^K \cosh h \pm \sqrt{e^{2K} \sinh^2 h + e^{-2K}} \nn\\
&\to \begin{cases}
2\cosh K \\
2\sinh K
\end{cases}\qquad (\textrm{when }h=0).
\ee
In the thermodynamic limit $M\to\oo$, the partition function is dominated by the larger eigenvalue
\be
Z_1 &=\l_+^M \p{1+\p{\frac{\l_-}{\l_+}}^M} \approx \l_+^M.
\ee
In quantum mechanical language, the state with the largest eigenvalue of $\hat T$ has the smallest eigenvalue of $\hat H$ --- i.e.\ it is the ground state, and we should call it $|0\>$. We have shown that the ground state dominates the thermodynamic limit. Contributions from the excited state are exponentially suppressed in the energy gap times the size of the system
\be
\p{\frac{\l_-}{\l_+}}^M &= e^{-(M\Delta\tau) m_\mathrm{gap}}, \nn\\
m_\mathrm{gap} &\equiv - \frac 1 {\Delta \tau} \log (\l_-/\l_+).
\ee
\stoplecture
We can also use the transfer matrix to compute correlation functions. For example, consider the two-point function $\<s_{i_1} s_{i_2}\>_{\Z_M}$ where the subscript $M$ indicates that we are on a periodic lattice with length $M$. Suppose $i_1>i_2$. We have
\be
\<s_{i_1} s_{i_2} \>_{\Z_M} &= \frac 1 {Z_1} \sum_{\{s_i\}} \<s_M|\hat{T}|s_{M-1}\> \cdots \<s_{i_1+1}|\hat{T}|s_{i_1}\> s_{i_1} \<s_{i_1}|\hat{T}|s_{i_1-1}\>\cdots \nn\\
& \qquad \x\cdots\<s_{i_2+1}|\hat{T}|s_{i_2}\> s_{i_2} \<s_{i_2}|\hat{T}|s_{i_2-1}\>\cdots \<s_1|\hat{T}|s_M\> \nn\\
&= \frac{1}{Z_1}\Tr(\hat{T}^{M-i_1} \hat{\s}^z \hat{T}^{i_1 - i_2}\s^z \hat T^{i_2}) \qquad (i_1 > i_2).
\label{eq:correlatorastrace}
\ee
Here, we introduced the Pauli spin operator $\s^z$ that measures the spin of a state
\be
\s^z|s\> &= s|s\>.
\ee
It is easy to compute the correlation function (\ref{eq:correlatorastrace}) by expressing $\hat s$ in the eigenbasis of $\hat T$.
\begin{exercise}
Show that in the limit of large $M$ and large ``distance" $i_1-i_2$, the correlator factorizes into a product of expectation values $\<0|\s^z|0\>$, plus exponential corrections from the excited state
\be
\<s_{i_1} s_{i_2}\>_{\Z_M} &= \<0|\s^z|0\>^2 + O(e^{-m_\mathrm{gap} \tau}, e^{- m_\mathrm{gap}(L-\tau)}),
\ee
where $\tau\equiv (i_1-i_2)\Delta\tau$, $L\equiv M\Delta\tau$.
\end{exercise}
Let us write (\ref{eq:correlatorastrace}) in a slightly different way by introducing ``Heisenberg picture" operators
\be
\hat s(i) &= \hat T^{-i} \s^z \hat T^i.
\ee
Equation~(\ref{eq:correlatorastrace}) is equivalent to
\be
\label{eq:heisenbergtwopt}
\<s_{i_1} s_{i_2}\>_{\Z_M} &= \frac 1 {Z_1} \Tr(\hat s(i_1) \hat s(i_2) \hat T^M) \qquad (i_1 > i_2).
\ee
Note that in deriving (\ref{eq:correlatorastrace}, \ref{eq:heisenbergtwopt}), we used that $i_1>i_2$. If instead $i_2 > i_1$, the path integral would give a product of operators in the opposite order
\be
\<s_{i_1} s_{i_2}\>_{\Z_M} &= \frac 1 {Z_1} \Tr(\hat s(i_2) \hat s(i_1) \hat T^M ) \qquad (i_2 > i_1).
\ee
The general statement is that the path integral becomes a time-ordered product of quantum operators\footnote{We hope that the time-ordering symbol $T\{\cdots\}$ will not be confused with the transfer matrix.}
\be
\label{eq:periodiccorrelationlattice}
\<s_{i_1} \cdots s_{i_n}\>_{\Z_M} &= \frac{1}{Z_1}\Tr(T\{ \hat s(i_1) \cdots \hat s(i_n)\}\hat T^M ),
\ee
where the definition of the time-ordering symbol is
\be
T\{\hat s (i_1)\cdots \hat s(i_n)\} &\equiv \hat s(i_1)\cdots \hat s(i_n) \theta(i_1 > \cdots > i_n) + \mathrm{permutations}.
\ee
Here $\theta(i_1 > \cdots > i_n)$ is $1$ if the $i_k$ are in the specified order and zero otherwise.
Finally, in the limit of large-$M$, the factor $\frac{\hat T^M}{\l_+^M}$ projects onto the ground state, so a path integral on the full lattice $\Z$ becomes a vacuum expectation value of a time-ordered product
\be
\label{eq:fulllattice}
\<s_{i_1}\cdots s_{i_n}\>_\Z &= \<0|T\{ \hat s(i_1) \cdots \hat s(i_n)\}|0\>.
\ee
\begin{exercise}
Consider adding a next-to-nearest neighbor interaction to the 1d Ising model
\be
S[s] &= \sum_{i} K s_{i}s_{i+1} + K' s_{i} s_{i+2}.
\ee
Find the associated quantum Hilbert space and transfer matrix, and compute the partition function on $\Z_M$ in the limit of large $M$.
\end{exercise}
\subsection{Quantization in quantum mechanics}
The procedure of turning a path integral into a product of quantum operators is called quantization. It is an extremely general procedure that ultimately stems from ``cutting and gluing" rules of the path integral.
For a more familiar example, let us review quantization for the path integral of a quantum particle on the real line. This theory has Euclidean action
\be
S[x] &= \int d\tau \p{\frac 1 2 \dot x^2 + V(x)},
\ee
where $x(\tau)$ is a map from Euclidean time $\tau$ to $\R$.
%To quantize, we introduce an orthonormal basis state $|x\>$ for every classical configuration at a fixed time $\tau_0$. These are the analogs of $|s=\pm 1\>$ in the Ising case, which were in one-to-one correspondence with classical configurations at a fixed lattice site $i_0$.
Consider the path integral on the interval $\tau\in[\tau_a,\tau_b]$ with fixed boundary conditions
\be
\label{eq:qmtransitionamplitude}
U(x_b,\tau_b;x_a,\tau_a) &\equiv \int_{\substack{x(\tau_b)=x_b \\ x(\tau_a) = x_a}} Dx(\tau) e^{-S[x]}.
\ee
By grouping paths according to their positions at a fixed time $\tau_c\in(\tau_a,\tau_b)$, we obtain a simple ``gluing" rule for $U$,
\be
\label{eq:gluingruleqm}
U(x_b,\tau_b;x_a,\tau_a) &= \int_{-\oo}^\oo dx_c U(x_b,\tau_b;x_c,\tau_c)U(x_c,\tau_c;x_a,\tau_a).
\ee
To quantize the theory, we build up the path integral in small time-increments using the gluing rule. Suppose we have already computed the path integral $U(x_0,\tau_0;x_a,\tau_a)$ on the interval $[\tau_a,\tau_0]$. To extend $\tau_0\to \tau_0+\e$, we write
\be
\label{eq:gluingforsmallinterval}
U(x_1,\tau_0+\e;x_a,\tau_a) &= \int dx_0\, U(x_1,\tau_0+\e;x_0,\tau_0) U(x_0,\tau_0;x_a,\tau_a).
\ee
The quantity
\be
U(x_1,\tau_0+\e;x_0,\tau_0) &= U(x_1,\e;x_0,0).
\ee
plays the role of a transfer matrix, and $U(x_0,\tau_0;x_a,\tau_a)$ plays the role of a state at time $\tau_0$. To see this, we introduce a Hilbert space
\be
\label{eq:qmhilberspace}
\mathcal{H} &= \mathrm{Span}\{|x\> : x\in \R\}
\ee
with inner product $\<x'|x\>=\de(x-x')$.
The $|x\>$ are analogs of the basis states $|s=\pm 1\>$ in the Ising model. Defining $\hat T_\e$ and $|\Psi(\tau_0)\>$ by
\be
\<x_1|\hat T_\e|x_0\> &= U(x_1,\e;x_0,0),\nn\\
\<x_0|\Psi(\tau_0)\> &= U(x_0,\tau_0;x_a,\tau_a),
\ee
equation (\ref{eq:gluingforsmallinterval}) becomes
\be
\label{eq:transfermatrixforqmeq}
|\Psi(\tau_0+\e)\> &= \hat T_\e |\Psi(\tau_0)\>.
\ee
To recover the Schrodinger equation, we must show that
\be
\hat T_\e &= 1 - \e \hat H + O(\e^2),
\ee
where
\be
\hat H &= -\frac 1 2 \frac{\ptl^2}{\ptl x^2} + V(x)
\ee
is the usual quantum-mechanical Hamiltonian. This is a standard argument, and we give a quick version here for completeness. We have
\be
\<x_1|\hat T_\e|x_0\> &= \int_{\substack{x(\e)=x_1 \\ x(0) = x_0}} Dx(\tau) e^{-S[x]},\qquad(\e \ll 1)\nn\\
S[x] &= \int_0^\e d\tau \p{\frac 1 2 \dot x^2 + V(x)}.
\ee
Because the time interval is so short, if $|x(\tau)-x_0|$ ever becomes larger than $O(\e)$, the kinetic term will cause the amplitude to be highly suppressed. Thus, let us assume $|x(\tau)-x_0|$ is of order $\e$. This means we can replace $V(x)\to V(x_0)$ (up to subleading corrections in $\e$), so the potential factors out of the integrand
\be
e^{-S[x]} &= e^{-\e V(x_0)} \exp\p{-\int_0^\e d\tau \frac{\dot x^2}{2}}.
\ee
We can now split $x(\tau)$ into a classical term and a fluctuation term $\z(\tau)$ with boundary conditions $\z(0)=\z(\e)=0$,
\be
x(\tau) &= x_0\p{1-\frac{\tau}{\e}} + x_1\frac{\tau}{\e} + \z(\tau),\nn\\
\int_0^\e d\tau \frac{\dot x^2}{2} &= \frac{(x_1-x_0)}{2\e} + \int_0^\e d\tau \frac{\dot\z^2}{2}.
\ee
The path integral over $\z$ contributes a constant $A_\e$ that depends on how the theory is regulated.\footnote{A very simple regulator is to approximate $x(\tau)$ as a piecewise linear path with segments of length $\e$. This is equivalent to simply setting $\z=0$ and not doing the integral.} We thus find
\be
\<x_1|\hat T|x_0\> &= A_\e e^{-\e V(x_0)} e^{-\frac{(x_1-x_0)^2}{2\e}}(1+O(\e)).
\ee
Finally, the Gaussian factor can be expanded in $\e$ using
\be
\label{eq:simplifyexponential}
e^{-\frac{x^2}{2\e}} &= \sqrt{2\pi \e}\p{\de(x) + \frac \e 2 \de''(x) + O(\e^2)},
\ee
which gives
\be
\<x_1|\hat T|x_0\> &= A_\e \sqrt{2\pi \e} \p{\de(x_1-x_0) - \e\p{-\frac 1 2\de''(x_1-x_0) + V(x_0)\de(x_1-x_0)} + O(\e^2)} \nn\\
&= A_\e \sqrt{2\pi \e}\<x_1|1 - \e \hat H + O(\e^2)|x_0\>.
\ee
The prefactor $A_\e \sqrt{2\pi \e}$ is an overall regulator-dependent constant that can be renormalized away by adding a ``cosmological constant" term to the action.
Thus, in the limit $\e\to 0$, equation (\ref{eq:transfermatrixforqmeq}) becomes a Euclidean version of the Schrodinger equation
\be
\frac{d}{d\tau} |\Psi(\tau)\> &= - \hat H|\Psi(\tau)\>.
\ee
Integrating using the initial condition $U(x',0,x,0)=\de(x-x')$, we find
\be
\label{eq:pathintegralwithbcsasamplitude}
U(x_b,\tau_b;x_a,\tau_a) &= \<x_b|e^{-\hat H (\tau_b-\tau_a)} |x_b\>.
\ee
Path integrals on different geometries correspond to different quantum mechanical observables. For example, the path integral on a circle $S^1_\b$ of length $\b$ with periodic boundary conditions is equal to
\be
\int_{-\oo}^\oo dx\, U(x,\b;x,0) = \Tr(e^{-\b \hat H}),
\ee
which is the quantum partition function at inverse temperature $\b$.
This is the continuum analog of equation~(\ref{eq:partitionfunctionastransfermatrix}).
Inserting observables into the path integral gives a time-ordered product of the associated quantum operators. For example, with periodic boundary conditions,
\be
\label{eq:periodiccorrelation}
\<x(\tau_1) \cdots x(\tau_n)\>_{S^1_\b} &=
\int_{x(\b)=x(0)} Dx\, x(\tau_1) \cdots x(\tau_n) e^{-S[x]}\nn\\
&= \Tr(T\{\hat x(\tau_1) \cdots \hat x(\tau_n)\} e^{-\b \hat H}),
\ee
where we introduced the operator
\be
\hat x(\tau) &= e^{\tau \hat H} \hat x e^{-\tau \hat H}, \nn\\
\hat x|x\> &= x|x\>.
\ee
To derive (\ref{eq:periodiccorrelation}), we cut the interval $[0,\b]$ into segments $[0,\tau_1]\cup [\tau_1,\tau_2] \cup \cdots [\tau_n,\b]$ (assuming $\tau_n > \cdots > \tau_1$) and use (\ref{eq:pathintegralwithbcsasamplitude}) for each segment.
The operator $\hat x$ is the analog of $\hat s$ and equation~(\ref{eq:periodiccorrelation}) is the analog of (\ref{eq:periodiccorrelationlattice}).
Finally, in the limit $\b\to \oo$, the factor $e^{-\b \hat H}$ projects onto the ground state. Thus the Euclidean path integral on $\R$ is equal to a time-ordered correlation function in the ground state,
\be
\<x(\tau_1) \cdots x(\tau_n)\>_{\R} &= \<0|T\{\hat x(\tau_1) \cdots \hat x(\tau_n)\}|0\>.
\ee
This is the continuum analog of (\ref{eq:fulllattice}).
\stoplecture
\subsection{The 2d Ising model}
Let us now consider a slightly more complicated case: the 2d Ising model. This will be a good toy example for how cutting, gluing, and quantization work in a general QFT. For simplicity, we set $h=0$. We consider the partition function on the doubly-periodic lattice $\Z_{M} \x \Z_{N}$ and label spins $s_{i,j}$ by a pair $(i,j) \in \Z_{M} \x \Z_{N}$.
The action is given by
\be
S[s] &= -K\sum_{i,j} (s_{i,j}s_{i+1,j} + s_{i,j}s_{i,j+1}) \nn\\
&= K \sum_{i,j} \p{\frac 1 2 (s_{i,j+1} - s_{i,j})^2 - 1} - K \sum_{i,j} s_{i,j} s_{i+1,j} \nn\\
&= \mathrm{const.} + \sum_{j=1}^{N} L( \bs_{j+1}, \bs_{j}).
\label{eq:unimportant}
\ee
In the last line, we split the action into contributions from pairs of neighboring rows.
The notation $s_{j}$ represents the configuration of spins in the $j$-th row,
\be
(\bs_{j})_{i} &= s_{i,j}.
\ee
The action associated with a pair of neighboring rows is given by
\be
L(\bs',\bs) &= \frac 1 2 K \sum_{i=1}^{M}(\bs'_i - \bs_i)^2 - \frac 1 2 K \sum_{i=1}^{M}(\bs_{i+1} \bs_i + \bs'_{i+1} \bs'_i).
\ee
(The constant in (\ref{eq:unimportant}) gives an unimportant multiplicative constant in the path integral that will disappear in normalized correlation functions, so we will ignore it.)
To quantize the theory, we can think of the $j$ direction as time, so $\bs_{j}$ is interpreted as a classical configuration on a fixed time-slice. The Hilbert space has an orthonormal basis vector for each such configuration,
\be
\label{eq:twodisinghilbert}
\mathcal{H}_{M} &= \mathrm{Span}\left\{|\pm\! 1, \pm 1, \cdots,\pm 1\>\right\} \nn\\
&= \bigotimes_{i=1}^{M} \mathcal{H}_i,
\ee
where $\mathcal{H}_i$ is a 1-qubit Hilbert space for each site $i$. $\mathcal{H}_{M}$ is the quantum Hilbert space of $M$ qubits, and is $2^{M}$-dimensional.
The transfer matrix between successive time slices is a $2^{M} \x 2^{M}$ matrix with entries
\be
\<\bs'|\hat T|\bs\> &= e^{-L(\bs',\bs)}.
\ee
The partition function on $\Z_{M} \x \Z_{N}$ is then
\be
\label{eq:partitionfunctionastracetwod}
Z(\Z_{M} \x \Z_{N}) &= \Tr_{\mathcal{H}_{M}}(\hat T^{N}).
\ee
To compute correlation functions, we need an operator that measures the spin at site $i$. This is simply the Pauli spin matrix $\s^z_i$ associated with the $i$-th site
\be
\s^z_i |s_1,\dots,s_j,\dots,s_{M}\> &= s_i |s_1,\dots,s_i,\dots,s_{M}\>.
\ee
We also have Heisenberg picture operators
\be
\label{eq:oldspinop}
\hat s_{i,j} &= \hat T^{-j} \s_i^z \hat T^j.
\ee
Correlation functions become traces of time-ordered products, e.g.
\be
\<s_{i_1,j_1} s_{i_2,j_2}\> &= \Tr_{\mathcal{H}_{M}}(\hat T^{N+j_2-j_1} \s_{i_1}^z \hat T_2^{j_1-j_2} \s_{i_2}^z) \th (j_1-j_2) + (1\leftrightarrow 2) \nn\\
&= \Tr_{\mathcal{H}_M}(T\{\hat s_{i_1,j_1} \hat s_{i_2,j_2}\} \hat T^{N}).
\ee
Let us now emphasize an important new ingredient in the 2-dimensional case compared to the 1-dimensional case. To arrive at (\ref{eq:partitionfunctionastracetwod}), we had to choose $j$ as the time direction. We then cut the path integral along rows of constant $j$. However, we could just as well have chosen $i$ as the time direction and cut the path integral along columns of constant $i$. This would give a different Hilbert space $\mathcal{H}_N$ with dimension $2^N$, a new transfer matrix $\hat T'$ (acting on $\mathcal{H}_N$), and a different formula for the same path integral,
\be
Z(\Z_M\x\Z_N) &= \Tr_{\mathcal{H}_N}(\hat T'^M) = \Tr_{\mathcal{H}_M}(\hat T^N).
\ee
In this new quantization, an insertion of $s_{i,j}$ in the path integral becomes an insertion of
\be
\label{eq:newspinop}
s_{i,j} &\to \hat T'^{-i} \s_j^z \hat T'^i
\ee
in a time-ordered product of quantum operators. Let us emphasize that the operators (\ref{eq:oldspinop}) and (\ref{eq:newspinop}) truly are different, even though they represent the same path integral variable. They even act on different-dimensional Hilbert spaces ($2^M$ vs.\ $2^N$)! Furthermore, even the notion of ``time"-ordering is different in this different quantization: in one case, operators are ordered according to their $i$ coordinates, and in the other they are ordered according to their $j$ coordinates.
\begin{exercise}
Show how to quantize the 2d Ising model in yet another way, using $i+j$ as the time coordinate and $i-j$ as space coordinate. What is the Hilbert space? What is the transfer matrix?
\end{exercise}
%More generally, we can build up the 2d Ising path integral by cutting and gluing path integrals on arbitrary regions on the lattice. For example, given a subset of spins $B\subset \Z^2$, let the ``boundary" $\ptl B$ consist of the spins in $B$ that have nontrivial interactions with spins outside of $B$. We use $s_b$ to indicate spins in the boundary $\ptl B$ and $s_i$ to indicate spins in the interior. We can associate a Hilbert space to $\ptl B$ by introducing an orthonormal basis vector for every spin configuration on $\ptl B$,
%\be
%\mathcal{H}_{\ptl B} &= \mathrm{Span}\left\{|s_b\> : \textrm{$s_b$ is a configuration of spins on $\ptl B$} \right\}.
%\ee
%The path integral over the interior with fixed boundary conditions gives a state $|\Psi\> \in \mathcal{H}$ with wavefunction
%\be
%\<s_{b 0}|\Psi\> &= \sum_{\substack {s_i=\pm 1 \\ s_b = s_{b0}}} e^{-S[s]}.
%\ee
%Consider now the complement $\bar B$, which also has $\ptl B$ as a boundary, and let $|\Psi'\>\in \mathcal{H}$ be the state produced by performing the path integral over the interior of $\bar B$. The path integral over $\bar B\cup B$ is given by taking a product of the associated wavefunctions and summing over boundary spins
%\be
%Z(\bar B\cup B) &= \sum_{s_b} \<\Psi'|s_b\>\<s_{b}|\Psi\> = \<\Psi'|\Psi\>.
%\ee
%
%In general when two regions $B_1,B_2$ share part of their boundary, we can partially glue states in $\mathcal{H}_{\ptl B_1}$ and $\mathcal{H}_{\ptl B_2}$ by summing over the shared boundary spins. This gives a state in the Hilbert space $\mathcal{H}_{\ptl(B_1\cup B_2)}$ (figure~\ref{}).
%
%In the special case where we cut the path integral into pieces related by a translation symmetry, we can introduce a transfer matrix and interpret correlators in terms of time-ordered products of quantum operators.
\subsection{Atiyah-Segal axioms}
\label{sec:asaxioms}
We are now ready to understand some axioms of continuum QFT. We will simultaneously state them and give examples in a Lagrangian theory of a real scalar field $\f$. A version of these axioms in TQFTs is due to Atiyah and a version in 2d CFTs is due to Segal. We will be interested in QFTs that depend on the geometry of the space in which they live, so when we say ``manifold" below, we mean ``Riemannian manifold with metric."
Consider a $d$-dimensional QFT $Q$.
\begin{enumerate}
\item To every $(d-1)$-manifold $N$ without boundary, $Q$ assigns a Hilbert space $\mathcal{H}_{N}$ called the space of states on $N$.
For example, in theories with an explicit path integral over some fields (``Lagrangian theories"), $\mathcal{H}_N$ has an orthonormal basis state for every field configuration on $N$. In scalar field theory,
\be
\label{eq:scalarfieldhilbertspace}
\mathcal{H}_N &= \mathrm{Span}\{|\f_b\> : \f_b\in C(N,\R)\}.
\ee
with inner product $\<\f_{b1}|\f_{b2}\> = \de(\f_{b1}-\f_{b2})$ (a functional $\de$-function).
Here $C(N,\R)$ is a space of real-valued functions on $N$.\footnote{We are being deliberately vague about {\it which\/} space of functions on $N$, e.g.\ $C^1(N)$, $C^\oo(N)$, etc.. In reality, the path integral is defined by introducing a regulator and taking a limit as the regulator is removed. Different regulators will result in different spaces of functions, and we prefer to be agnostic about which regulator is being used.} More formally, we can write
\be
\cH_N &= L^2(C(N,\R),\C),
\ee
where $L^2(X)$ denotes the square-integrable complex functions on $X$.
The space $\cH_N$ is the analog of the Hilbert spaces (\ref{eq:qmhilberspace}) and (\ref{eq:twodisinghilbert}) for quantum mechanics and the 2d Ising model. Let us discuss the example of quantum mechanics more explicitly. Quantum mechanics is a $1$-dimensional QFT. In this case $d-1=0$, and the only possible $0$-manifolds are disjoint unions of points. The space of field configurations on a point is $C(\bullet,\R)=\R$, so the Hilbert space associated to a point is $\cH_\bullet = L^2(C(\bullet,\R),\C) = L^2(\R,\C)$. This is spanned by the usual basis states $|x\>$ with inner product $\de(x-x')$.\footnote{Note in particular that $L^2(C(\bullet,\R),\C)$ is totally different from $L^2(\bullet,\C)=\C$. In general, don't confuse the Hilbert space $\cH_N$ with the space of square-integrable functions $L^2(N,\C)$.}
%Note that $L^2(C(N,\R),\C)$ totally different from $L^2(N,\C)$. To understand this, consider the special case of quantum mechanics, which is a $1$-dimensional QFT. The only possible $(d-1)$-manifolds are then disjoint unions of points. The space of fields on a point is $C(\bullet,\R)=\R$, so the Hilbert space is $\cH_\bullet = L^2(C(\bullet,\R)) = L^2(\R,\C)$. This is spanned by the usual basis states $|x\>$ with inner product $\de(x-x')$. This is not the same as $L^2(\bullet,\C)=\C$.
\item The space of states on a disconnected union is a tensor product
\be
\label{eq:disjointunionrule}
\mathcal{H}_{N_1 \sqcup N_2} &= \mathcal{H}_{N_1}\otimes \mathcal{H}_{N_2}.
\ee
This should be clear in scalar theory from the definition (\ref{eq:scalarfieldhilbertspace}). It will be useful to assign a Hilbert space to the empty manifold, and the only reasonable choice consistent with (\ref{eq:disjointunionrule}) is
\be
\mathcal{H}_{\varnothing} &= \mathbb{C}.
\ee
\item To every $d$-manifold $M$ with incoming boundary $N$ and outgoing boundary $N'$, $Q$ assigns a transition amplitude\footnote{By incoming vs. outgoing boundary, we assume that each boundary component $N$ comes equipped with a nonzero infinitesimal vector field pointing into or out of $M$.}
\be
\label{eq:transitionamplitude}
Z_M : \mathcal{H}_N \to \mathcal{H}_{N'}.
\ee
In Lagrangian theories, $Z_M$ is obtained by fixing boundary conditions on $N,N'$ and performing the path integral over the interior. For example in a scalar theory, the matrix elements of $Z_M$ are given by
\be
\label{eq:transitionmatrixelement}
\<\f_b'|Z_M|\f_{b}\> &= \int_{\substack{\phi|_{N} = \f_b \\ \phi|_{N'} = \f_b'}} D\phi(x) e^{-S[\f]}.
\ee
Here, the notation $\f|_{N}$ means the restriction of $\f$ to $N$.
In the special case where $M$ is a closed manifold (without boundary), $Z_M$ is a linear map
\be
\label{eq:closedmanifold}
Z_M : \mathbb{C} \to \mathbb{C}.
\ee
A linear map $\C\to\C$ is simply multiplication by complex number. That number is the ``partition function" of the QFT on $M$.
Another important case is where $M$ has boundaries $(\varnothing,N)$. (The first element of the ordered pair is the incoming boundary and the second is the outgoing boundary.) In this case, $Z_M$ gives a linear map
\be
\label{eq:outgoingboundary}
Z_M: \C \to \mathcal{H}_N.
\ee
Any such map is uniquely determined by its action on $1\in \C$, so this is equivalent to a state $|\Psi\>=Z_M(1) \in \mathcal{H}_N$. Thus, we can ``prepare a state" by performing the path integral with fixed boundary conditions. We saw several examples in the previous sections, for example the definition of $Z_\mathrm{partial}(i,s)$ in the 1d Ising model.
Finally, consider the case where $M$ has boundaries $(N,\varnothing)$. Then we have a linear map
\be
\label{eq:incomingboundary}
Z_M: \mathcal{H}_N \to \C,
\ee
which is equivalent to a bra-state $\<\Psi|\in \mathcal{H}_N^*$.
The usefulness of assigning $\mathcal{H}_\varnothing=\C$ is that all these special cases (\ref{eq:closedmanifold}, \ref{eq:outgoingboundary}, \ref{eq:incomingboundary}) are covered by one axiom (\ref{eq:transitionamplitude}).
\item Suppose $M_1$ has boundaries $(N,N')$ and $M_2$ has boundaries $(N',N'')$. We can form a new $d$-manifold $M_1\cup_{N'} M_2$ by gluing $M_1$ and $M_2$ along $N'$. This manifold has boundaries $(N,N'')$. The total transition amplitude is the composition
\be
\label{eq:compositionrule}
Z_{M_1\cup_{N'} M_2} &= Z_{M_2} Z_{M_1} : \mathcal{H}_N \to \mathcal{H}_{N''}.
\ee
In scalar field theory, this follows by cutting the path integral along $N'$ by fixing boundary conditions $\f_b'\in C(N',\R)$, and then finally performing the $(d-1)$-dimensional path integral over $\f_b'$,
\be
&\int_{\substack{\phi|_{N} = \f_b \\ \phi|_{N''} = \f_b''}} D\phi(x) e^{-S[\f]} \nn\\
&= \int_{y\in N'} D\f_b'(y)
\int_{\substack{\phi_2|_{N'} = \f_b' \\ \phi_1|_{N''} = \f_b'' \\ x \in M_2}} D\phi_2(x) e^{-S[\f_2]}
\int_{\substack{\phi_1|_{N} = \f_b \\ \phi_1|_{N'} = \f_b' \\ x \in M_1}} D\phi_1(x) e^{-S[\f_1]}.
\ee
This is the field-theory analog of our gluing rule for transition amplitudes in quantum mechanics (\ref{eq:gluingruleqm}).
In the notation (\ref{eq:transitionmatrixelement}), we have simply inserted a complete set of states on $N'$,
\be
\<\f_b''|Z_{M_2} Z_{M_1}|\f_b\> &= \int_{y\in N'} D\f'_b(y) \<\f_b''|Z_{M_2}|\f_b'\>\<\f_b'|Z_{M_1}|\f_b\>.
\ee
\item \draftnote{Describe generalization to correlation functions.}
\item \draftnote{Discuss orientation, orientation reversal, conjugation.}
\end{enumerate}
\draftnote{Discuss the possibility of giving the manifolds different amounts of structure, for instance topological structure, a Riemannian metric, etc. In these notes, we will initially be interested in manifolds with a metric. Later, we will see that in certain cases we can relax this structure to only a conformal class of metrics.}
Although we have given explicit prescriptions in the case of a continuum theory, all of these axioms have analogs in the case of the 2d Ising model. It is not hard to imagine how the continuum axioms might arise in the continuum limit of the Ising model, at least in the case where the ``manifolds" $M,N,\dots$ are regions of flat space.
\stoplecture
\subsection{Quantization in continuum QFT}
Let us use the above axioms to describe the quantization procedure one more time, now in continuum QFT.
Consider the $d$-manifold $M_L=I_{L} \x N$, where $I_L=[0,L]$ is an interval of length $L$ and $N$ is a $(d-1)$-manifold, and suppose $M_L$ is endowed with the product metric. The composition rule (\ref{eq:compositionrule}) implies that
\be
\label{eq:exponentiatedtransition}
Z_{M_L} &= e^{-L \hat H}
\ee
for some operator
\be
\hat H:\mathcal{H}_N \to \mathcal{H}_N,
\ee
namely the Hamiltonian.
Consider a correlation function on $S_\b^1 \x N$
\be
\label{eq:correlatorcontinuum}
\<\f(\tau_1,y_1)\cdots \f(\tau_n,y_n)\>_{S_\b^1 \x N} &=
\int_{\f(\b,y)=\f(0,y)} D\f(x)\, \f(\tau_1,y_1)\cdots \f(\tau_n,y_n) e^{-S[\f]}.
\ee
Here, $\tau,y$ are coordinates on $S_\b^1,N$, respectively.
As before, we introduce operators $\hat \f(y)$ that measure the field configuration on a time-slice
\be
\hat \f(y)|\f_b\> &= \f_b(y)|\f_b\>.
\ee
To compute (\ref{eq:correlatorcontinuum}), we cut the path integral at times $\tau_1,\dots,\tau_n$ and repeatedly use (\ref{eq:exponentiatedtransition}). By the usual logic, this gives
\be
\<\f(\tau_1,y_1)\cdots \f(\tau_n,y_n)\>_{S_\b^1 \x N} &= \Tr_{\mathcal{H}_N}(e^{-\b\hat H} T\{\hat \f(\tau_1,y_1)\cdots \hat \f(\tau_n,y_n)\}),
\ee
where
\be
\hat \f(\tau,y) &\equiv e^{\tau \hat H} \hat \f(y) e^{-\tau \hat H},
\ee
and $T\{\cdots\}$ represents time ordering in $\tau$.
Taking the limit of a long interval, the factor $e^{-\b \hat H}$ projects onto the ground state, and we obtain\footnote{In (\ref{eq:limitoflongbeta}), we are assuming that the theory has a unique ground state on $N$. More generally, there might be multiple ground states on $N$, in which case the limit $\b\to \oo$ becomes a trace in the space of ground states.}
\be
\label{eq:limitoflongbeta}
\<\f(\tau_1,y_1)\cdots \f(\tau_n,y_n)\>_{\R \x N} &=\<0|T\{\hat \f(\tau_1,y_1)\cdots \hat \f(\tau_n,y_n)\}|0\>.
\ee
For a QFT in $\R^d$, we can quantize the theory in many different ways by choosing different directions to play the role of ``time." In every quantization of the theory, we have
\be
\<\f(x_1)\cdots \f(x_n)\>_{\R^d} &= \<0|T\{\hat \f(x_1)\cdots \hat \f(x_n)\}|0\>.
\ee
However, in different quantizations, the objects appearing on the right-hand side are different. The Hilbert spaces are different (though isomorphic if the theory is $\SO(d)$-invariant), the ground states $|0\>$ are different, the quantum operators $\hat\f(x)$ are different, and the time-ordering symbols means different things. However, because they originate from a single path integral, the resulting expectation values are the same. It is sometimes useful to think of expressions like the RHS above as ``interpretations" of one underlying path integral.
\draftnote{Discuss boundary conditions and whether we should think of $\R^d$ as having a boundary or not.}
\subsection{The quantum transverse-field Ising model}
After all this abstract nonsense, let's return to more concrete things and try to understand the critical point of the 2d Ising model. The transfer matrix $\hat T$ of the 2d Ising lattice model was diagonalized in 1944 by Onsager. Unfortunately, we won't have time to describe his solution. Instead, we will study a closely related model in the same universality class as the lattice model.
Recall that the transfer matrix has matrix elements $\<\bs'|\hat T|\bs\> = e^{-L(\bs',\bs)}$. Let us write $\hat T$ in a more familiar way as an operator on a spin chain. First split $L$ into contributions from horizontal and vertical bonds
\be
L(\bs',\bs) &= L_h(\bs')+L_h(\bs) + L_v(\bs',\bs). \nn\\
L_h(\bs) &= -\frac 1 2 K \sum_i \bs_{i+1} \bs_i, \nn\\
L_v(\bs',\bs) &= \sum_i \frac 1 2 K(\bs_i'-\bs_i)^2.
\ee
Note that
\be
e^{\frac 1 2 K \sum_i \s^z_i \s^z_{i+1}}|\bs\> &= e^{-L_h(\bs)}|\bs\>.
\ee
Meanwhile, $L_v$ only involves spins at a single site, so let us imagine that we have only one site. Note that
\be
\<s'|(1+e^{-2K}\s^x)|s\> &= e^{-\frac 1 2 K (s'-s)^2}.
\ee
We also have
\be
1+e^{-2K} \s^x &= e^{A+K' \s^x},\qquad\textrm{where} \nn\\
\tanh K' &= e^{-2K},\nn\\
e^A &= \sqrt{1-e^{-4K}},
\ee
which follows by expanding out the Taylor series for $e^{A+K'\s^x}$ and matching the coefficients of $1,\s^x$.
Thus,