forked from ribault/CFT-Review
-
Notifications
You must be signed in to change notification settings - Fork 0
/
crash.tex
703 lines (610 loc) · 35.6 KB
/
crash.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
% Let us try to further condense the 'minimal lectures' in order to give a 2h course at Cargèse 2017.
% I managed to do Section 1 in one hour, but the rest did not fit in one hour.
\documentclass[12pt, a4paper]{article}
\raggedbottom
\RequirePackage[l2tabu, orthodox]{nag}
\usepackage[top=20mm,bottom=20mm,left=25mm,right=25mm]{geometry}
% \usepackage{showkeys}
\usepackage{amsfonts, amsmath}
\usepackage{multirow}
\usepackage{pgf}
\usepackage{tikz}
\usepackage{amsthm}
\newtheoremstyle{break}{9pt}{9pt}{\itshape}{}{\bfseries}{}{\newline}{}
\theoremstyle{break}
\newtheorem{exo}{Exercise}
\newtheorem{hyp}[exo]{Axiom}
\newtheorem{res}[exo]{Result}
\newtheorem{defn}[exo]{Definition}
\usepackage[colorlinks=true,linktoc=all,linkcolor=black,citecolor=red,urlcolor=blue]{hyperref}
\title{\bfseries A crash course on two-dimensional \\ conformal field theory}
\author{Sylvain Ribault \vspace{2mm}
\\
{\normalsize CEA Saclay, Institut de Physique Th\'eorique}
\\
{\footnotesize \ttfamily sylvain.ribault@ipht.fr }
}
\begin{document}
\maketitle
\begin{abstract}
We provide a brief but self-contained introduction to conformal field theory on the Riemann sphere. We first introduce general axioms such as local conformal invariance, and derive Ward identities and BPZ equations. We then define Liouville theory by specific axioms on its spectrum and degenerate fields. We solve the theory and study its four-point functions.
\end{abstract}
\tableofcontents
%\vspace{10mm}
%\noindent \textit{Systematically announce plan and axiom numbers!}
\subsection*{Acknowledgements}
I am grateful to Samuel Belliard, Rouven Frassek, and Beno\^it Vallet for valuable suggestions.
\clearpage
\hypersetup{linkcolor=blue}
%\numberwithin{equation}{section}
This text was written with brevity as the main concern, so as to be the basis for about 100 minutes' worth of lectures. For more details, see the ``Minimal lectures on 2d CFT'' \cite{rib16} (20 pages) and the full review article \cite{rib14} (120 pages), both on arXiv. For fewer details, see the Wikipedia articles on ''Two-dimensional conformal field theory'' and ''Liouville theory'', that I mostly wrote.
The first two sections are about the main ideas of 2d CFT, as developed in the 1980s. Then we will focus on the particular case of Liouville theory, a CFT that was mostly solved in the 1990s, with however some subtleties understood only in 2015. Liouville theory was introduced by Polyakov as a model of 2d gravity, and later applied to string theory. More recently, it has been related to statistical systems such as Conformal Loop Ensembles (Ikhlef--Jacobsen--Saleur) or Random Energy Models (Raoul Santachiara's talks in this school).
\section{Conformal field theory}
\subsection{The Virasoro algebra and its representations}
By definition, conformal transformations are transformations that preserve angles.
In two dimensions with a complex coordinate $z$, any holomorphic map preserves angles, and
infinitesimal conformal transformations have the generators
\begin{align}
\ell_n = -z^{n+1}\frac{\partial}{\partial z}\qquad (n\in\mathbb{Z})\ ,
\end{align}
including $\ell_{-1}=-\frac{\partial}{\partial z}$ for translations.
The commutation relations are
\begin{align}
[\ell_n,\ell_m ] = (n-m)\ell_{m+n}\ ,
\end{align}
and the corresponding Lie algebra is called the Witt algebra.
In a quantum theory, symmetry transformations act projectively on states.
Projective representations of an algebra are equivalent to representations of a centrally extended algebra.
\begin{defn}[Virasoro algebra]
~\label{def:vir}
The central extension of the Witt algebra is called the Virasoro algebra. It has the generators $(L_n)_{n\in\mathbb{Z}}$ and $\mathbf 1$, and the commutation relations
\begin{align}
[\mathbf 1, L_n] = 0 \quad , \quad [L_n,L_m] = (n-m)L_{n+m} +\frac{c}{12}(n-1)n(n+1)\delta_{n+m,0}\mathbf 1 \ ,
\label{eq:vir}
\end{align}
where the number $c$ is called the central charge.
\end{defn}
The spectrum, i.e. the space of states, must be a representation of the Virasoro algebra.
We want to interpret $L_0$ as the energy operator, so its eigenvalues must be bounded from below. But
\begin{align}
L_0 |v\rangle = \Delta |v\rangle \qquad \implies \qquad L_0 L_n|v\rangle = (\Delta-n)L_n|v\rangle \ .
\end{align}
So $L_{n>0}$ decrease conformal dimensions, and must therefore kill the state with the lowest dimension.
\begin{defn}[Primary and descendent states, level, Verma module]
~\label{def:prim}
A primary state with conformal dimension $\Delta$ is a state $|v\rangle$ such that
\begin{align}
L_0 |v\rangle = \Delta |v\rangle \quad , \quad L_{n>0} |v\rangle = 0\ .
\end{align}
The Verma module $\mathcal V_\Delta$ is the representation whose basis is
$
\left\{ \prod_{i=1}^k L_{-n_i} |v\rangle\right\}_{ 0<n_1\leq \dots \leq n_k}
$.
The level of the state $\prod_{i=1}^k L_{-n_i} |v\rangle $ is $N=\sum_{i=1}^k n_i\geq 0$. A state of level $N\geq 1$ is called a descendent state.
\end{defn}
Let us plot a basis of a Verma module up to the level $N=2$:
\begin{align}
\begin{tikzpicture}[scale = .25, baseline=(current bounding box.center)]
\draw[-latex, very thick] (20, 0) -- (20, -14) node [right] {$N$};
\foreach \x in {0, ..., 2}
{
\draw [dotted] (-20, {-6*\x}) -- (20, {-6*\x}) node [right] {${\x}$};
}
\node[fill = white] at (0, 0) (0) {$|v\rangle$};
\node[fill = white] at (-4,-6) (1) {$L_{-1}|v\rangle$};
\node[fill = white] at (-8, -12) (11) {$L_{-1}^2|v\rangle$};
\node[fill = white] at (4,-12) (2) {$L_{-2}|v\rangle$};
\draw[-latex] (0) -- (1);
\draw[-latex] (1) -- (11);
\draw[-latex] (0) -- (2);
\end{tikzpicture}
\end{align}
If a Verma module is not irreducible, i.e. if it has a nontrivial submodule, then the state with the lowest conformal dimension in that submodule must again be a primary state, while also being a descendent state.
\begin{defn}[Null vectors]
~\label{def:nv}
A descendent state that is also primary is called a null vector or singular vector.
\end{defn}
In the Verma module $\mathcal V_\Delta$, let us find out whether $L_{-1}|v\rangle$ is a null vector:
\begin{align}
L_n L_{-1}|v\rangle \underset{n\geq 1}{=} [L_n, L_{-1}] |v\rangle = (n+1) L_{n-1}|v\rangle =
\left\{\begin{array}{ll} 0 & \quad \text{if } n\geq 2 \\ 2\Delta |v\rangle & \quad \text{if } n = 1 \end{array}\right.
\end{align}
So $L_{-1}|v\rangle$ is a null vector if and only if $\Delta=0$, and the Verma module $\mathcal V_0$ is reducible.
Similarly, there exists a level two null vector if and only if
\begin{align}
\Delta = \frac{1}{16}\left( 5-c\pm\sqrt{(c-1)(c-25)} \right) \ .
\label{eq:dpm}
\end{align}
In order to simplify this formula, let us introduce other notations for $c$ and $\Delta$. We define
\begin{align}
\text{the background charge } Q \ , & \quad c = 1+6Q^2\ ,
\\
\text{the coupling constant } b \ , & \quad Q = b+\frac{1}{b} \ ,
\\
\text{the momentum } \alpha\ , &\quad \Delta = \alpha(Q-\alpha)\ .
\end{align}
The condition \eqref{eq:dpm} for the existence of a level two singular vector becomes
\begin{align}
\alpha = -\frac12 b^{\pm 1}\ .
\end{align}
To summarize, the null vectors $L|v\rangle$ at levels $N\leq 2$ and the conditions on $\alpha$ for their existence are
\begin{align}
\renewcommand{\arraystretch}{1.3}
\begin{array}{|c|c|c|c|}
\hline
N & \alpha_{\langle r,s\rangle} & L_{\langle r,s\rangle} & \langle r,s\rangle
\\
\hline\hline
1 & 0 & L_{-1} & \langle 1,1\rangle
\\
\hline
\multirow{2}{*}{2} & -\frac{b}{2} & L_{-1}^2 + b^2 L_{-2} &
\langle 2,1\rangle
\\
\cline{2-4}
& -\frac{1}{2b} & L_{-1}^2 + b^{-2} L_{-2} & \langle 1,2\rangle
\\
\hline
\end{array}
\label{lot}
\end{align}
The general pattern is that null vectors at a level $N$ are parametrized by pairs $r,s$ of strictly positive integers such that $rs=N$.
For a null vector to exist, the momentum must be of the type
\begin{align}
\alpha_{\langle r,s\rangle} = \frac{Q}{2} - \frac12(rb+sb^{-1})\qquad \text{with} \qquad N=rs\ .
\label{eq:ars}
\end{align}
For generic $c$, the submodule generated by a null vector is the Verma module $\mathcal V_{\Delta_{\langle r,s\rangle} +rs}\subset \mathcal V_{\Delta_{\langle r,s\rangle}}$. Then the coset $\frac{ \mathcal V_{\Delta_{\langle r,s\rangle}} }{ \mathcal V_{\Delta_{\langle r,s\rangle} +rs} }$ is an irreducible representation, where the null vector vanishes.
\subsection{Fields and correlation functions}
Let us define fields on the Riemann sphere $\mathbb{C}\cup\{\infty\}$.
\begin{hyp}[State-field correspondence]
~\label{hyp:sfc}
For any state $|w\rangle$ in the spectrum, there is an associated field $V_{|w\rangle}(z)$. The map $|w\rangle \mapsto V_{|w\rangle}(z)$ is linear. We define the action of the Virasoro algebra on such fields as
\begin{align}
L_n V_{|w\rangle}(z) = L_n^{(z)} V_{|w\rangle}(z) = V_{L_n|w\rangle}(z)\ .
\end{align}
\end{hyp}
\begin{defn}[Primary field, descendent field, degenerate field]
~\label{def:pfdf}
Let $|v\rangle$ be the primary state of the Verma module $\mathcal V_\Delta$.
We define the primary field $V_\Delta(z)=V_{|v\rangle}(z)$. This field obeys
\begin{align}
L_{n\geq 0} V_\Delta(z) = 0 \quad , \quad L_0 V_\Delta(z) = \Delta V_\Delta(z)\ .
\end{align}
Similarly, descendent fields correspond to descendent states. And the degenerate field $V_{\langle r,s\rangle}(z)$ corresponds to the primary state of the coset representation $\frac{ \mathcal V_{\Delta_{\langle r,s\rangle}} }{ \mathcal V_{\Delta_{\langle r,s\rangle} +rs} }$, and therefore obeys
\begin{align}
L_{\langle r, s\rangle} V_{\langle r,s\rangle}(z) = 0\qquad \text{-- for example, \ } L_{-1}V_{\langle 1,1\rangle}(z) = 0\ .
\end{align}
\end{defn}
\begin{hyp}[Dependence of fields on $z$]
~\label{hyp:geom}
For any field $V(z)$, we have
\begin{align}
\frac{\partial}{\partial z} V(z) = L_{-1} V(z) \ .
\label{pvlv}
\end{align}
\end{hyp}
Using this axiom for both $V(z)$ and $L_n^{(z)}V(z)$, we find
\begin{align}
\frac{\partial}{\partial z} L_n^{(z)} = [L_{-1}^{(z)},L_n^{(z)}]= -(n+1)L_{n-1}^{(z)}\ ,\qquad (\forall n\in\mathbb{Z})\ .
\end{align}
These infinitely many equations can be encoded into one functional equation,
\begin{align}
\frac{\partial}{\partial z} \sum_{n\in\mathbb{Z}} \frac{L_n^{(z)}}{(y-z)^{n+2}} = 0\ .
\end{align}
\begin{defn}[Energy-momentum tensor]
~\label{def:em}
The energy-momentum tensor is a field, that we define by the formal Laurent series
\begin{align}
T(y) = \sum_{n\in\mathbb{Z}} \frac{L_n^{(z)}}{(y-z)^{n+2}} \ .
\end{align}
In other words, for any field $V(z)$, we have
\begin{align}
T(y)V(z) = \sum_{n\in\mathbb{Z}} \frac{L_n V(z)}{(y-z)^{n+2}}\ .
\label{eq:lvtv}
\end{align}
\end{defn}
In particular, for a primary field $V_\Delta(z)$, we find
\begin{align}
T(y)V_\Delta(z) = \frac{\Delta}{(y-z)^2} V_\Delta(z) + \frac{1}{y-z} \frac{\partial}{\partial z} V_\Delta(z) + O(1)\ .
\label{eq:tvd}
\end{align}
This is our first example of an operator product expansion.
The energy-momentum tensor $T(y)$ is locally holomorphic as a function of $y$, and acquires poles in the presence of other fields. Since we are on the Riemann sphere, it must also be holomorphic at $y=\infty$.
\begin{hyp}[Behaviour of $T(y)$ at infinity]
~\label{hyp:ti}
\begin{align}
T(y) \underset{y\to\infty} = O\left(\frac{1}{y^4}\right)\ .
\label{eq:tinf}
\end{align}
\end{hyp}
\begin{defn}[Correlation function]
~\label{def:cor}
To $N$ fields $V_1(z_1), \dots ,V_N(z_N)$ with $z_i\neq z_j$, we associate a number called their correlation function or $N$-point function, and denoted as
\begin{align}
\Big< V_1(z_1) \cdots V_N(z_N) \Big>\ .
\end{align}
Correlation functions depend linearly on fields.
\end{defn}
\begin{hyp}[Commutativity of fields]
~\label{hyp:ass}
Correlation functions do not depend on the order of the fields,
\begin{align}
V_1(z_1) V_2(z_2) = V_2(z_2)V_1(z_1)\ .
\end{align}
\end{hyp}
\subsection{Ward identities and BPZ equations}
Let us work out the consequences of conformal symmetry for correlation functions.
In order to study an $N$-point function $Z$ of primary fields, we introduce an auxiliary $(N+1)$-point function $Z(y)$ where we insert the energy-momentum tensor,
\begin{align}
Z = \left< \prod_{i=1}^N V_{\Delta_i}(z_i) \right> \quad , \quad Z(y) = \left< T(y) \prod_{i=1}^N V_{\Delta_i}(z_i) \right> \ .
\end{align}
$Z(y)$ is a meromorphic function of $y$, with poles at $y=z_i$, whose residues are given by eq. \eqref{eq:tvd} (using the commutativity of fields). This implies
\begin{align}
Z(y) = \sum_{i=1}^N \left(\frac{\Delta_i}{(y-z_i)^2} +\frac{1}{y-z_i}\frac{\partial}{\partial z_i}\right) Z\ .
\label{eq:zy}
\end{align}
Moreover, the behaviour \eqref{eq:tinf} of $T(y)$ implies that
the coefficients of $y^{-1}, y^{-2}, y^{-3}$ in the large $y$ expansion of $Z(y)$ must vanish,
\begin{align}
\sum_{i=1}^N \partial_{z_i} Z = \sum_{i=1}^N \left(z_i \partial_{z_i} + \Delta_i\right) Z = \sum_{i=1}^N \left(z_i^2 \partial_{z_i} + 2\Delta_iz_i\right) Z = 0\ .
\label{eq:gward}
\end{align}
These three equations are called global Ward identities.
Let us solve them. For a one-point function, we have
\begin{align}
\partial_z \Big< V_\Delta(z)\Big> = 0 \ , \quad \Delta \Big< V_\Delta(z)\Big> = 0\ , \quad \implies \quad \Big< V_\Delta(z)\Big> \propto \delta_{\Delta, 0}\ .
\end{align}
In the case of two-point functions, we find
\begin{align}
\Big< V_{\Delta_1}(z_1)V_{\Delta_2}(z_2) \Big> \propto \delta_{\Delta_1,\Delta_2} (z_1-z_2)^{-2\Delta_1} \ .
\label{eq:2pt}
\end{align}
For three-point functions, there are as many equations \eqref{eq:gward} as unknowns $z_1,z_2,z_3$, and therefore a unique solution with no constraints on $\Delta_i$,
\begin{align}
\left< \prod_{i=1}^3 V_{\Delta_i}(z_i) \right> \propto (z_1-z_2)^{\Delta_3-\Delta_1-\Delta_2}
\times \text{$2$ permutations}\ .
\label{eq:3pt}
\end{align}
For four-point functions, the general solution is
\begin{align}
\left< \prod_{i=1}^4 V_{\Delta_i}(z_i) \right> = \left(\prod_{i<j}(z_i-z_j)^{\delta_{ij}}\right) G\left(\frac{(z_1-z_2)(z_3-z_4)}{(z_1-z_3)(z_2-z_4)}\right)\ ,
\end{align}
where $\sum_{j\neq i} \delta_{ij} = -2\Delta_i$, and $G$ is an arbitrary function.
Equivalently, the three global Ward identities effectively reduces the four-point function to a function of one variable,
\begin{align}
G(z) = \Big< V_{\Delta_1}(z) V_{\Delta_2}(0)V_{\Delta_3}(\infty)V_{\Delta_4}(1) \Big>\ .
\end{align}
Correlation functions of descendents can be deduced from correlation functions of primaries. For example,
\begin{align}
\Big< L_{-2}V_{\Delta_1}(z_1) V_{\Delta_2}(z_2)\cdots \Big>
= \frac{1}{2\pi i}\oint_{z_1} \frac{dy}{y-z_1} Z(y)
=\sum_{i=2}^N\left(\frac{1}{z_1-z_i}\frac{\partial}{\partial z_i} +\frac{\Delta_i}{(z_i-z_1)^2}\right) Z\ ,
\label{eq:ltv}
\end{align}
where we used first eq. \eqref{eq:lvtv} for $L_{-2}V_{\Delta_1}(z_1)$, and then eq. \eqref{eq:zy} for $Z(y)$.
This can be generalized to any correlation function of descendent fields. The resulting equations are called local Ward identities.
Correlation functions that involve degenerate fields obey additional equations. In particular,
\begin{align}
\left(L_{-1}^2 + b^2 L_{-2}\right) V_{\langle 2, 1 \rangle}(z_1) = 0\qquad \text{so that} \qquad L_{-2}V_{\langle 2, 1 \rangle}(z_1) = -\frac{1}{b^2}\frac{\partial^2}{\partial z_1^2} V_{\langle 2, 1 \rangle}(z_1)\ .
\end{align}
Together with the local Ward identity \eqref{eq:ltv},
this leads to the second-order Belavin--Polyakov--Zamolodchikov partial differential equation
\begin{align}
\left( \frac{1}{b^2}\frac{\partial^2}{\partial z_1^2} + \sum_{i=2}^N\left(\frac{1}{z_1-z_i}\frac{\partial}{\partial z_i} +\frac{\Delta_i}{(z_1-z_i)^2}\right) \right)\left< V_{\langle 2, 1 \rangle}(z_1) \prod_{i=2}^N V_{\Delta_i}(z_i) \right> = 0\ .
\label{eq:bpz}
\end{align}
In the case of a four-point function $N=4$, the BPZ equation amounts to a differential equation for a function of one variable.
\section{Conformal bootstrap}
Conformal symmetry leads to linear equations for correlation functions: Ward identities and BPZ equations.
In order to fully determine correlation functions, we need additional, nonlinear equations, and therefore additional axioms: single-valuedness of correlation functions, and existence of operator product expansions.
Using these axioms for studying conformal field theories is called the conformal bootstrap method.
\subsection{Single-valuedness}
\begin{hyp}[Single-valuedness]
~\label{hyp:sv}
Correlation functions are single-valued functions of the positions, i.e. they have trivial monodromies.
\end{hyp}
Our two-point function \eqref{eq:2pt} however has nontrivial monodromy unless $\Delta_1\in \frac12\mathbb{Z}$, as a result of solving holomorphic Ward identities.
We would rather have a single-valued function of the type $|z_1-z_2|^{-4\Delta_1} = (z_1-z_2)^{-2\Delta_1} (\bar z_1-\bar z_2)^{-2\Delta_1}$.
This suggests that we need anti-holomorphic Ward identities as well, and therefore a second copy of the Virasoro algebra.
\begin{hyp}[Left and right Virasoro algebras]
~\label{hyp:lr}
We have two mutually commuting Virasoro symmetry algebras with the same central charge, called left-moving or holomorphic, and right-moving or anti-holomorphic. Their generators are written $L_n,\bar L_n$, with in particular
\begin{align}
\frac{\partial}{\partial z} V(z) = L_{-1}V(z) \quad , \quad \frac{\partial}{\partial \bar z} V(z)= \bar L_{-1} V(z) \ .
\end{align}
\end{hyp}
Let us consider left- and right-primary fields $V_{\Delta_i,\bar\Delta_i}(z_i)$, with the
two-point functions
\begin{align}
\left<\prod_{i=1}^2 V_{\Delta_i,\bar\Delta_i}(z_i) \right> \propto \delta_{\Delta_1,\Delta_2}\delta_{\bar\Delta_1,\bar\Delta_2} (z_1-z_2)^{-2\Delta_1} (\bar z_1-\bar z_2)^{-2\bar\Delta_1}\ .
\end{align}
This is single-valued if and only if our two fields have half-integer spins,
\begin{align}
\Delta -\bar \Delta \in \frac12\mathbb{Z}\ .
\end{align}
The simplest case is $\Delta=\bar\Delta$, which leads to the definition
\begin{defn}[Diagonal states, diagonal fields and diagonal spectrums]
~\label{def:diag}
A primary state or field is called diagonal if it has the same left and right conformal dimensions. A spectrum is called diagonal if all primary states are diagonal.
\end{defn}
From now on we assume that all primary fields are diagonal, and write $V_\Delta(z) = V_{\Delta,\Delta}(z)$.
\subsection{Operator product expansion and crossing symmetry}
\begin{hyp}[Operator product expansion]
~\label{hyp:ope}
Let $V_1(z_1)$ and $V_2(z_2)$ be two fields, and $|w_i\rangle$ be a basis of the spectrum.
There exist coefficients $C^i_{12}(z_1,z_2)$ such that we have the operator product expansion (OPE)
\begin{align}
V_1(z_1)V_2(z_2) = \sum_i C^i_{12}(z_1,z_2) V_{|w_i\rangle}(z_2)\ .
\end{align}
\end{hyp}
If the spectrum is made of diagonal primary states and their descendent states, the OPE of two primary fields must be of the type
\begin{align}
V_{\Delta_1}(z_1) V_{\Delta_2}(z_2)
= \sum_{\Delta\in S} C_{\Delta_1,\Delta_2,\Delta} |z_1-z_2|^{2(\Delta-\Delta_1-\Delta_2)}
\Big(V_{\Delta}(z_2) + O(z_1-z_2) \Big)\ ,
\label{eq:ope}
\end{align}
where $C_{\Delta_1,\Delta_2,\Delta}$ is the $z_i$-independent OPE coefficient.
$O(z_1-z_2)$ denotes the contributions of descendents fields, which are determined by conformal Ward identities.
Using the OPE, we can reduce a three-point function to a combination of two-point functions, and we find
\begin{align}
\left<\prod_{i=1}^3 V_{\Delta_i}(z_i) \right> = C_{\Delta_1,\Delta_2,\Delta_3} |z_1-z_2|^{2(\Delta_3-\Delta_1-\Delta_2)} |z_1-z_3|^{2(\Delta_2-\Delta_1-\Delta_3)} |z_2-z_3|^{2(\Delta_1-\Delta_2-\Delta_3)}\ ,
\end{align}
so that the OPE coefficient $C$ coincides with the three-point structure constant.
Inserting the OPE in a four-point function of primary fields, we find
\begin{align}
\Big<V_{\Delta_1}(z)V_{\Delta_2}(0)V_{\Delta_3}(\infty)V_{\Delta_4}(1)\Big>
=\sum_{\Delta\in S} C_{\Delta_1,\Delta_2,\Delta} C_{\Delta,\Delta_3,\Delta_4} \mathcal{F}^{(s)}_\Delta(z) \mathcal{F}^{(s)}_\Delta(\bar z)\ .
\label{sdec}
\end{align}
\begin{defn}[Conformal block]
~\label{def:block}
The four-point conformal block on the sphere,
\begin{align}
\mathcal{F}^{(s)}_\Delta(z) \underset{z\to 0}{=} z^{\Delta-\Delta_1-\Delta_2}\Big( 1 + O(z) \Big)\ ,
\label{eq:gsd}
\end{align}
is the normalized contribution of the Verma module $\mathcal V_\Delta$ to a four-point function, obtained by summing over left-moving descendents. It is a locally holomorphic function of $z$. Its dependence on $c,\Delta_1,\Delta_2,\Delta_3,\Delta_4$ are kept implicit. The label $(s)$ stands for for $s$-channel.
\end{defn}
Conformal blocks are in principle known, as they are universal functions, entirely determined by conformal symmetry. This is analogous to characters of representations, also known as zero-point conformal blocks on the torus.
Our axiom \ref{hyp:ass} on the commutativity of fields implies that the OPE is associative, and that we can use the OPE of any two fields in a four-point function. In particular, using the OPE of the first and fourth fields, we obtain
\begin{align}
\Big<V_{\Delta_1}(z)V_{\Delta_2}(0)V_{\Delta_3}(\infty)V_{\Delta_4}(1)\Big>
=\sum_{\Delta\in S} C_{\Delta,\Delta_1,\Delta_4} C_{\Delta_2,\Delta_3,\Delta} \mathcal{F}^{(t)}_\Delta(z) \mathcal{F}^{(t)}_\Delta(\bar z)\ ,
\label{tdec}
\end{align}
where $\mathcal{F}^{(t)}_\Delta(z) \underset{z\to 1}{=} (z-1)^{\Delta-\Delta_1-\Delta_4}\Big(1+O(z-1)\Big)$ is a $t$-channel conformal block.
The equality of our two decompositions \eqref{sdec} and \eqref{tdec} of the four-point function is called crossing symmetry, schematically
\begin{align}
\sum_{s} C_{12s} C_{s34}
\begin{tikzpicture}[baseline=(current bounding box.center), very thick, scale = .3]
\draw (-1,2) node [left] {$2$} -- (0,0) -- node [above] {$s$} (4,0) -- (5,2) node [right] {$3$};
\draw (-1,-2) node [left] {$1$} -- (0,0);
\draw (4,0) -- (5,-2) node [right] {$4$};
\end{tikzpicture}
= \sum_{t} C_{23t}C_{t41}
\begin{tikzpicture}[baseline=(current bounding box.center), very thick, scale = .3]
\draw (-2,3) node [left] {$2$} -- (0,2) -- node [left] {$t$} (0,-2) -- (-2, -3) node [left] {$1$};
\draw (0,2) -- (2,3) node [right] {$3$};
\draw (0,-2) -- (2, -3) node [right] {$4$};
\end{tikzpicture}
\ .
\label{csd}
\end{align}
Given the spectrum $S$, crossing symmetry is a system of quadratic equations for the structure constant $C_{\Delta_1,\Delta_2,\Delta_3}$. Requiring that this system has solutions is a strong constraint on the spectrum.
\begin{defn}[Conformal field theory]
~\label{def:cft}
A (model of) conformal field theory on the Riemann sphere is a spectrum $S$ and a set of correlation functions $\left<\prod_{i=1}^N V_{|w_i\rangle}(z_i)\right>$ with $|w_i\rangle\in S$ that obey all our axioms, in particular crossing symmetry.
\end{defn}
\begin{defn}[Defining and solving]
~\label{def:def}
To define a conformal field theory is to give principles that uniquely determine its spectrum and correlation functions.
To solve a conformal field theory is to actually compute them.
\end{defn}
\subsection{Hypergeometric four-point functions}
Crossing symmetry equations typically involve infinite sums, which makes them difficult to solve.
However, if at least one field is degenerate, then the four-point function belongs to the finite-dimensional space of solutions of a BPZ equation.
For example,
$G(z)= \Big< V_{\langle 2, 1 \rangle}(z) V_{\Delta_1}(0)V_{\Delta_2}(\infty)V_{\Delta_3}(1) \Big>$ is a combination of only two $s$-channel conformal blocks.
These blocks are a particular basis of solutions of the BPZ equation, characterized by their asymptotic behaviour near $z=0$ \eqref{eq:gsd},
\begin{align}
\mathcal{F}^{(s)}_{\alpha_1-\frac{b}{2}}(z) = z^{b\alpha_1} (1-z)^{b\alpha_3} F(A,B,C,z)\quad , \quad
\mathcal{F}^{(s)}_{\alpha_1+\frac{b}{2}}(z) = \left. \mathcal{F}^{(s)}_{\alpha_1-\frac{b}{2}}(z) \right|_{\alpha_1\to Q-\alpha_1} \ ,
\label{gpm}
\end{align}
where $F(A,B,C,z)$ is the hypergeometric function with parameters
\begin{align}
\renewcommand{\arraystretch}{1.3}
\left\{\begin{array}{l} A = \frac12 + b(\alpha_1+\alpha_3-Q) + b(\alpha_2-\tfrac{Q}{2}) \ , \\
B = \frac12 + b(\alpha_1+\alpha_3-Q) - b(\alpha_2-\tfrac{Q}{2}) \ , \\
C = 1 + b(2\alpha_1-Q) \ .
\end{array}\right.
\label{abc}
\end{align}
Let us build single-valued four-point functions as linear combinations of such blocks. Single-valuedness near $z=0$ forbids terms such as $\mathcal{F}^{(s)}_{\alpha_1-\frac{b}{2}}(z) \mathcal{F}^{(s)}_{\alpha_1+\frac{b}{2}}(\bar z)$, and we must have
\begin{align}
G(z) = \sum_{i=\pm} c^{(s)}_{i} \mathcal{F}^{(s)}_{\alpha_1-i\frac{b}{2}}(z) \mathcal{F}^{(s)}_{\alpha_1-i\frac{b}{2}}(\bar z) = \sum_{j=\pm} c^{(t)}_{i} \mathcal{F}^{(t)}_{\alpha_3-j\frac{b}{2}}(z) \mathcal{F}^{(t)}_{\alpha_3-j\frac{b}{2}}(\bar z)\ .
\end{align}
The $s$- and $t$-channel blocks are two bases of the same space of solutions of the BPZ equation, and they are linearly related,
\begin{align}
\mathcal{F}^{(s)}_{\alpha_1-i\frac{b}{2}}(z) = \sum_{j=\pm} F_{ij} \mathcal{F}^{(t)}_{\alpha_3-j\frac{b}{2}}(z)\ .
\end{align}
In particular, this implies
\begin{align}
\frac{c_{+}^{(s)}}{c_{-}^{(s)}} = -\frac{F_{-+}F_{--}}{F_{++}F_{+-}}
= \frac{\gamma(A)\gamma(B)\gamma(C-A)\gamma(C-B)}{\gamma(C)\gamma(C-1)}\quad \text{with} \quad \gamma(x) =\frac{\Gamma(x)}{\Gamma(1-x)}\ .
\label{eq:coc}
\end{align}
We will later express $c_\pm^{(s)}$ in terms of three-point structure constants, and obtain equations for these structure constants.
\section{Liouville theory}
\subsection{Definition}
\begin{defn}[Liouville theory]
~\label{def:liou}
For any value of the central charge $c\in\mathbb{C}$, Liouville theory is the conformal field theory whose space of states is
\begin{align}
\int_{\frac{Q}{2}+i\mathbb{R}_+} d\alpha\ \mathcal V_\alpha \otimes
\bar{\mathcal V}_\alpha\ ,
\end{align}
and whose correlation functions are analytic functions of $b$ and $\alpha$, assuming it exists and its unique.
\end{defn}
Why these values of $\alpha$? If $c\in \mathbb{R}$ we want $\Delta\in\mathbb{R}$ i.e. $\alpha\in (\frac{Q}{2}+i\mathbb{R}) \cup \mathbb{R}$. But we also want $\Delta$ to be bounded from below, and the natural lower bound is $\Delta(\alpha=\frac{Q}{2}) =\frac{c-1}{24}$.
Anyway, the most important part of the definition is the assumption of analyticity in $\alpha$.
The integral is actually over the half-line $\frac{Q}{2}+i\mathbb{R}_+$ so as not to count representations twice.
We indeed have $\Delta(\alpha)=\Delta(Q-\alpha)$ so $\mathcal{V}_\alpha = \mathcal{V}_{Q-\alpha}$.
The fields $V_\alpha(z)$ and $V_{Q-\alpha}(z)$ then correspond to the same primary state, and we must have a reflection relation,
\begin{align}
V_\alpha(z) = R(\alpha) V_{Q-\alpha}(z)\ ,
\end{align}
where the function $R(\alpha)$ is called the reflection coefficient.
Let us schematically write three-point functions and OPEs in Liouville theory,
\begin{align}
%\Big< V_{\alpha_1}V_{\alpha_2} \Big> &= \delta(Q-\alpha_1-\alpha_2) + R(\alpha_1)\delta(\alpha_1-\alpha_2)\ ,
%\label{eq:vv}
%\\
\Big< V_{\alpha_1}V_{\alpha_2}V_{\alpha_3} \Big> & = C_{\alpha_1,\alpha_2,\alpha_3}
%\quad \text{with} \quad C_{\alpha_1,\alpha_2,\alpha_3} = R(\alpha_1) C_{Q-\alpha_1,\alpha_2,\alpha_3}
\ ,
\\
V_{\alpha_1}V_{\alpha_2} &= \int_{\frac{Q}{2}+i\mathbb{R}_+} d\alpha\ C_{\alpha_1,\alpha_2,Q-\alpha} \Big( V_\alpha + \cdots\Big)\ .
\label{eq:v1v2}
\end{align}
In order to have reasonably simple crossing symmetry equations, we need degenerate fields. But the spectrum is made of Verma modules, and the corresponding fields are not degenerate. So we need a special axiom.
\begin{hyp}[Degenerate fields in Liouville theory]
~\label{hyp:degl}
The degenerate fields $V_{\langle 2, 1\rangle}$ and $V_{\langle 1, 2\rangle}$, and their correlation functions, exist.
\end{hyp}
Let us introduce notations for OPE coefficients of $V_{\langle 2, 1\rangle}$:
\begin{align}
V_{\langle 2, 1\rangle} V_\alpha \sim C_+(\alpha) V_{\alpha-\frac{b}{2}} + C_-(\alpha)V_{\alpha +\frac{b}{2}}\ .
\end{align}
It can be shown that there exists a field renormalization $V_\alpha(z) \to \lambda(\alpha)V_\alpha(z)$ such that $C_+(\alpha)=1$.
Then, using the reflection relation, we can deduce
\begin{align}
C_+(\alpha) = 1\quad , \quad C_-(\alpha) = \frac{R(\alpha)}{R(\alpha+\frac{b}{2})}\ .
\end{align}
\subsection{Three-point structure constants}
In the case of the four-point function $\Big\langle V_{\langle 2,1 \rangle}(z) V_\alpha(0) V_{Q-\alpha}(\infty) V_{\langle 2,1 \rangle}(1)\Big\rangle$, we have
\begin{align}
c_{+}^{(s)} = C_+(\alpha)C_-(\alpha-\tfrac{b}{2}) = \frac{R(\alpha-\frac{b}{2})}{R(\alpha)} \quad , \quad c_{-}^{(s)} = C_-(\alpha)C_+(\alpha+\tfrac{b}{2}) = \frac{R(\alpha)}{R(\alpha+\frac{b}{2})}\ ,
\end{align}
and eq. \eqref{eq:coc} boils down to
\begin{align}
\frac{R(\alpha-\frac{b}{2})R(\alpha+\frac{b}{2})}{R(\alpha)^2}
= \frac{\gamma(2b\alpha)\gamma(b(Q-2\alpha))}{\gamma(b(2Q-2\alpha))\gamma(b(2\alpha-Q))}\ .
\end{align}
We could get the same equation with $b\to \frac{1}{b}$ by having $V_{\langle 1,2\rangle}$ instead of $V_{\langle 2,1\rangle}$ in the four-point function. The analytic solution of these equations is
\begin{align}
R(\alpha) = -b^2 \frac{\gamma(b(2\alpha-Q))}{\gamma(\frac{1}{b}(Q-2\alpha))}\ .
\end{align}
This also determines $C_\pm(\alpha)$.
In a more general hypergeometric four-point function the coefficients $c^{(s)}_\pm$ are
\begin{align}
c_{\pm}^{(s)} = C_\pm(\alpha_1) C_{\alpha_1\mp\frac{b}{2},\alpha_2,\alpha_3} \ .
\label{cs}
\end{align}
Using eq. \eqref{eq:coc}, we obtain the shift equation
\begin{align}
\frac{C_{\alpha_1+b,\alpha_2,\alpha_3}}{C_{\alpha_1,\alpha_2,\alpha_3}} = b^{-4}\frac{\gamma(2b\alpha_1)\gamma(2b\alpha_1+b^2)}{\prod_{\pm,\pm} \gamma\left(b\alpha_1\pm b(\alpha_2-\frac{Q}{2})\pm b(\alpha_3-\frac{Q}{2})\right)}\ .
\label{fcc}
\end{align}
Again, the equation with $b\to \frac{1}{b}$ should hold too. In order to solve these equations, we need a function that produces Gamma functions when its argument is shifted by $b$ or $\frac{1}{b}$.
\begin{defn}[Upsilon function]
~\label{def:upsilon}
For $b>0$, let $\Upsilon_b(x)$ be the unique (up to a constant factor) holomorphic function that obeys the shift equations
\begin{align}
\frac{\Upsilon_b(x+b)}{\Upsilon_b(x)} = b^{1-2bx} \gamma(bx)\qquad \text{and} \qquad \frac{\Upsilon_b(x+\frac{1}{b})}{\Upsilon_b(x)} = b^{\frac{2x}{b}-1} \gamma(\tfrac{x}{b})\ .
\label{upup}
\end{align}
For $ib>0$, the meromorphic function
\begin{align}
\hat{\Upsilon}_b(x) = \frac{1}{\Upsilon_{ib}(-ix+ib)}\ ,
\end{align}
then obeys the same shift equations.
\end{defn}
The functions $\Upsilon_b(x)$ and $\hat\Upsilon_b(x)$ can respectively be defined for $\Re b>0$ and $\Re ib>0$ by analytic continuation. And we have
\begin{align}
\Upsilon_b(x) = \prod_{m,n=0}^\infty f\left(\frac{\frac{Q}{2}-x}{\frac{Q}{2}+mb+nb^{-1}}\right) \quad \text{with} \quad f(x)=(1-x^2)e^{x^2}\ .
\end{align}
Using the functions $\Upsilon_b(x)$, we can write a solution $C$ of the shift equations \eqref{fcc} for three-point structure constants,
\begin{align}
C_{\alpha_1,\alpha_2,\alpha_3} = \frac{ \Upsilon_b(2\alpha_1) \Upsilon_b(2\alpha_2) \Upsilon_b(2\alpha_3)}{\Upsilon_b(\alpha_1+\alpha_2+\alpha_3-Q) \Upsilon_b(\alpha_1+\alpha_2-\alpha_3)\Upsilon_b(\alpha_2+\alpha_3-\alpha_1)\Upsilon_b(\alpha_3+\alpha_1-\alpha_2)} \ .
\label{caaa}
\end{align}
This solution is called the DOZZ formula for Dorn, Otto, A.
Zamolodchikov and Al. Zamolodchikov (1995). It holds if
$c\notin ]-\infty, 1]$ i.e. $b\notin i\mathbb{R}$.
On the other hand, doing the replacement $\Upsilon_b\to \hat\Upsilon_b$, we obtain a solution $\hat C$ that holds if $c\notin [25,\infty[$ i.e. $b\notin \mathbb{R}$ (2000s).
Using the analyticity of correlation functions,
the solution is unique if $b$ and $b^{-1}$ are aligned, i.e. if $b^2\in\mathbb{R}$:
\begin{equation}
\begin{tikzpicture}[baseline=(current bounding box.center), scale = .6]
\draw (0, 2) node[left]{$i$} -- (0, 1) node[below left] {$0$} -- (1, 1) node[below] {$1$};
\draw [thick, latex-latex] (11,3) -- (11,1) node[fill, circle, minimum size = 1mm, inner sep = 0]{} -- (11,-.3);
\draw [thick, latex-latex] (8,3) -- (7,1) node[fill, circle, minimum size = 1mm, inner sep = 0]{}-- (7.6,-.2);
\draw [thick, latex-latex] (5,1) -- (3,1) node[fill, circle, minimum size = 1mm, inner sep = 0]{} -- (4.3,1) ;
\node at (4, -1.5){$\begin{array}{c} b\in \mathbb{R} \\ c\geq 25 \end{array}$};
\node at (7.5, -1.5){$\begin{array}{c} b\in \mathbb{C} \\ c\in\mathbb{C} \end{array}$};
\node at (11, -1.5){$\begin{array}{c} b\in i\mathbb{R} \\ c\leq 1 \end{array}$};
\end{tikzpicture}
\end{equation}
However, for general values of $c$, both $C$ and $\hat C$ are solutions. In order to prove the existence and uniqueness of Liouville theory, we have to determine which solutions lead to crossing-symmetric four-point functions.
\subsection{Four-point functions}
The $s$-channel decomposition of a Liouville four-point function reads
\begin{align}
\Big< V_{\alpha_1}(z) V_{\alpha_2}(0) V_{\alpha_3}(\infty) V_{\alpha_4}(1)\Big> = \frac12 \int_{\frac{Q}{2}+i\mathbb{R}} d\alpha\ C_{\alpha_1,\alpha_2,Q-\alpha} C_{\alpha,\alpha_3, \alpha_4} \mathcal{F}_{\alpha}^{(s)}(z) \mathcal{F}_{\alpha}^{(s)}(\bar z)\ ,
\label{eq:ccff}
\end{align}
where the structure can be $C$ or $\hat C$. Let us accept that the conformal blocks $\mathcal{F}_{\alpha}^{(s)}(z)$ have poles when $\alpha = \alpha_{\langle r, s\rangle} $ \eqref{eq:ars}, the momentums for which the $s$-channel representation becomes reducible.
We now plot the positions of these poles (blue regions) relative to the integration line (red), depending on the central charge:
\begin{align}
\newcommand{\polewedge}[3]{
\begin{scope}[#1]
\node[blue, draw,circle,inner sep=1pt,fill] at (0, 0) {};
\node[#3] at (0,0) {#2};
\filldraw[opacity = .1, blue] (0,0) -- (4, -4) -- (4, 4) -- cycle;
\end{scope}
}
\begin{array}{ccc}
\begin{tikzpicture}[scale = .4, baseline=(current bounding box.center)]
\draw[ultra thick, blue, opacity = .3] (0,0) -- (-4,0);
\draw[ultra thick, blue, opacity = .3] (2,0) -- (6,0);
\draw[-latex] (-4,0) -- (0, 0) node[above left] {$0$} -- (6,0) node [above] {$\alpha$};
\draw (0, -4.5) -- (0, 4.5);
\draw[ultra thick, red] (1, -4.5) -- (1, 4.5);
\node[blue, draw,circle,inner sep=1pt,fill] at (0, 0) {};
\node[above left] at (0,0) {$0$};
\node[blue, draw,circle,inner sep=1pt,fill] at (2, 0) {};
\node[above right] at (2,0) {$Q$};
\end{tikzpicture}
&
\begin{tikzpicture}[scale = .4, baseline=(current bounding box.center)]
\draw[-latex] (-4,0) -- (0, 0) node[above left] {$0$} -- (6,0) node [above] {$\alpha$};
\draw (0, -4.5) -- (0, 4.5);
\draw[ultra thick, red] (1, -4.5) -- (1, 4.5);
\polewedge{rotate = 180}{$0$}{above left};
\polewedge{shift = {(2, .3)}}{$Q$}{above right};
\end{tikzpicture}
&
\begin{tikzpicture}[scale = .4, baseline=(current bounding box.center)]
\draw[-latex] (-3,0) -- (0, 0) node[above left] {$0$} -- (4,0) node [above] {$\alpha$};
\draw[ultra thick, blue, opacity = .3] (0, -4.5) -- (0, 4.5);
\draw (0, -4.5) -- (0, 4.5);
\draw[ultra thick, red] (.5, -4.5) -- (.5, 4.5);
\node[blue, draw,circle,inner sep=1pt,fill] at (0, 0) {};
\node[above left] at (0,0) {$0$};
\node[blue, draw,circle,inner sep=1pt,fill] at (0, 1.5) {};
\node[above left] at (0,1.5) {$Q$};
\end{tikzpicture}
\vspace{3mm}
\\
c\geq 25 & c\text{ generic} & c\leq 1
\end{array}
\end{align}
The four-point function built from $C$ is analytic on $c\notin ]-\infty,1]$. So if Liouville theory exists for $c\geq 25$, then it also exists for $c$ generic, with the same structure constant $C$.
On the other hand, the limit $c\to ]-\infty, 1]$ is singular. Actually, for $c\leq 1$, the integration line has to be slightly shifted in order to avoid the poles. So the structure constant $\hat C$ is expected to be valid only for $c\leq 1$.
That is how far we can easily get with analytic considerations.
Numerical tests of crossing symmetry confirm that Liouville theory exists for all values of $c$, with the three-point structure constants $\hat C$ if $c\leq 1$, and $C$ otherwise \cite{rs15} (2015).
\bibliographystyle{cft}
\bibliography{cft}
\end{document}