-
Notifications
You must be signed in to change notification settings - Fork 8
/
solvable.tex
4161 lines (3767 loc) · 321 KB
/
solvable.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
\documentclass[12pt, a4paper]{article}
\raggedbottom
\RequirePackage[l2tabu, orthodox]{nag}
\usepackage{silence}
\WarningFilter[pdftoc]{hyperref}{Token not allowed in a PDF string}
\ActivateWarningFilters[pdftoc]
\usepackage[top=18mm,bottom=18mm,left=25mm,right=25mm]{geometry}
\usepackage{showkeys}
\usepackage{amsmath,amssymb,amsfonts}
\usepackage[table]{xcolor}
\usepackage{centernot}
%\usepackage[nottoc]{tocbibind}
\newcommand{\aline}{\hspace{.5cm}\hfil\rule{0.93\textwidth}{.4pt}}
\usepackage{makeidx}
\makeindex
\usepackage[nottoc]{tocbibind}
\newcommand{\myindex}[1]{\textbf{\boldmath #1}}
\usepackage{pgf}
\usepackage{tikz}
\usetikzlibrary{shapes}
\usetikzlibrary{shapes.misc}
\tikzset{cross/.style={cross out, draw=black, thick, fill=none, minimum size=2*(#1-\pgflinewidth), inner sep=0pt, outer sep=0pt}, cross/.default={3pt}}
\newcommand{\uvertex}{
\coordinate (base) at (0, 1);
\draw (0, 0) node[fill, circle, minimum size = 1.3mm, inner sep = 0]{};
\draw (3, 0) node [cross]{};
\draw (0, 3) node[fill, circle, minimum size = 1.3mm, inner sep = 0]{};
\draw (3, 3) node[fill, circle, minimum size = 1.3mm, inner sep = 0]{};
}
\newcommand{\qvertex}{
\coordinate (base) at (0, 1);
\draw (0, 0) node [cross]{};
\draw (3, 0) node [cross]{};
\draw (0, 3) node[cross]{};
\draw (3, 3) node[cross]{};
}
\newcommand{\vertices}{
\coordinate (base) at (0, 1);
\draw (0, 0) node[fill, circle, minimum size = 1.3mm, inner sep = 0]{};
\draw (3, 0) node[fill, circle, minimum size = 1.3mm, inner sep = 0]{};
\draw (0, 3) node[fill, circle, minimum size = 1.3mm, inner sep = 0]{};
\draw (3, 3) node[fill, circle, minimum size = 1.3mm, inner sep = 0]{};
}
\usepackage{multirow}
\usepackage{xcolor}
% Syntax: \colorboxed[<color model>]{<color specification>}{<math formula>}
\newcommand*{\colorboxed}{}
\def\colorboxed#1#{%
\colorboxedAux{#1}%
}
\newcommand*{\colorboxedAux}[3]{%
% #1: optional argument for color model
% #2: color specification
% #3: formula
\begingroup
\colorlet{cb@saved}{.}%
\color#1{#2}%
\boxed{%
\color{cb@saved}%
#3%
}%
\endgroup
}
\usepackage[most]{tcolorbox}
\tcbset{
frame code={}
center title,
left=0pt,
right=0pt,
top=0pt,
bottom=0pt,
colback=green!15,
colframe=white,
% width=\dimexpr\textwidth\relax,
enlarge left by=0mm,
% boxsep=5pt,
arc=0pt,outer arc=0pt,
breakable,
}
\usepackage{amsthm}
\usepackage{thmtools}
%\usepackage{thm-restate} This package does not interact very smoothly with \theoremstyle and \listoftheorems
\newtheoremstyle{break}{9pt}{9pt}{\itshape}{}{\bfseries}{}{\newline}{}
\theoremstyle{break}
\newtheorem{exo}{Exercise}[section]
\newtheorem{hyp}[exo]{Axiom}
\newtheorem{res}[exo]{Result}
\newtheorem{defn}[exo]{Definition}
\makeatletter
\def\ll@exo{%
\protect\numberline{\csname the\thmt@envname\endcsname}%
\ifx\@empty\thmt@shortoptarg
\thmt@thmname
\else
\thmt@shortoptarg
\fi}
\def\l@thmt@exo{}
\makeatother
\usepackage[colorlinks=true,linktoc=all,linkcolor=black,citecolor=red,urlcolor=blue,backref=page]{hyperref}
\usepackage{etoolbox}
\makeatletter
\patchcmd{\BR@backref}{\newblock}{\newblock(}{}{}
\patchcmd{\BR@backref}{\par}{)\par}{}{}
\makeatother
\renewcommand{\arraystretch}{1.5}
\title{\bfseries Exactly solvable conformal field theories}
\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 review 2d CFT in the bootstrap approach, and sketch the known exactly solvable CFTs with no extended chiral symmetry: Liouville theory, (generalized) minimal models, limits thereof, and loop CFTs, including the $O(n)$, Potts and $U(n)$ CFTs.
Exact solvability relies on local conformal symmetry, and on the existence of degenerate fields. We show how these assumptions constrain the spectrum and correlation functions. We discuss how crossing symmetry equations can be solved analytically and/or numerically, leading to analytic expressions for structure constants in terms of the double Gamma function.
% In particular, we review the Ward identities and Virasoro blocks that follow from local conformal symmetry. And we formalize the existence of degenerate fields in terms of an interchiral algebra and interchiral blocks.
% Exact solvability also relies on analyticity assumptions for the dependence of correlation functions on the central charge and on the conformal dimensions of diagonal fields. In particular, we show how analyticity assumptions determine the logarithmic conformal blocks that appear in loop CFTs.
% Also mention generic complex central charge? analyticity whenever continuous variables.
% We discuss how crossing symmetry equations can be solved analytically or numerically.
% Analytic solutions lead to shift equations for structure constants, which can be solved in terms of the double Gamma function. The numerical approach is needed when we have infinitely many linear equations for infinitely many unknowns: we review how to determine the space of solutions, and particular solutions.
In the case of loop CFTs, we sketch the corresponding statistical models, and derive the relation between statistical and CFT variables. We review the resulting combinatorial description of correlation functions, and discuss what remains to be done for solving the CFTs.
\end{abstract}
% Last sentence not justified in text. Add subdivision to 5.5? Would be more or less recap of article... talk about backbone? already done elsewhere.
\clearpage
\tableofcontents
\hypersetup{linkcolor=blue}
\numberwithin{equation}{section}
%\setcounter{secnumdepth}{2}
%\setcounter{tocdepth}{2}
\setcounter{section}{-1}
\pagebreak
\section{Introduction}
\subsubsection*{Exact solvability in CFT}
In 1997, di Francesco, Mathieu and Sénéchal wrote a book called Conformal field theory \cite{fms97}, which dealt almost exclusively with conformal field theory \textit{in two dimensions}. Today, a book with that title would be expected to also cover CFT in $d\geq 3$. However, when it comes to CFTs that we can solve exactly, we are back to $d=2$, because exact solvability relies on 2 ingredients:
\begin{itemize}
\item An infinite-dimensional algebra of conformal transformations.
\item Degenerate fields: fields whose OPEs with other fields have finitely many terms, modulo conformal symmetry.
\end{itemize}
This gives us access to powerful bootstrap techniques, analytic and numerical. But even in $d=2$, solvable CFTs are very special cases. Starting from a simple CFT and following a renormalization group flow, we can easily end up at a fixed point whose CFT is intractable.
In $d\geq 3$, the conformal algebra is finite-dimensional. Degenerate fields can however be constructed, under the name of weight-shifting operators \cite{kks17}. Under an OPE with a weight-shifting operator, the conformal dimension of a field is shifted by an integer. But except in particular CFTs such as generalized free theories, the spectrum is not stable under such shifts. So weight-shifting operators cannot be consistenly added to the CFT, and do not constrain it, although they are still useful for studying conformal blocks. The main bootstrap techniques in $d\geq 3$ are numerical instead, and rely on the assumption of unitarity. Not only this restricts the CFTs that can be addressed, but also this leads to inequalities rather than equalities, making it harder to reach exact results.
\subsubsection*{Solvable CFTs}
We will first deal with CFTs that are considered solved, in that we exactly know the 3-point structure constants, and can compute any correlation function to arbitrary precision:
\begin{itemize}
\item Liouville theory,
\item (generalized) minimal models, and limits thereof.
\end{itemize}
We will also propose a unified treatment of loop CFTs, which are not solved but are believed to be solvable, because we have sufficiently many exact results. By loop CFTs we mean a large class of CFTs, with various global symmetry groups, which describe critical limits of various statistical models in various phases. The known solvable CFTs have many applications to statistical physics, in particular to percolation, to polymers, and more recently to active hydraulics \cite{jcpb23}. Liouville theory and minimal models have well-known applications to quantum gravity in 2d. Applications of non-diagonal solvable CFTs such as loop CFTs to quantum gravity in 3d are limited by the bound on the central charge $\Re c < 13$. (See Section \ref{sec:dotc}.) A CFT that would be holographically dual to quantum gravity in 3d Anti-de Sitter space would need to exist for $c\to+\infty$, while for 3d de Sitter space we would need $\Re c = 13$ \cite{god24, cemr24}.
There are other solved or solvable CFTs that we will not mention, because they have extended chiral algebras, which imply extra local symmetries in addition to conformal symmetry. Some CFTs of that type can be solved using the methods that we will cover, for example certain Wess--Zumino--Witten models.
\subsubsection*{Solving CFTs}
When a CFT is solvable, it can typically be solved in several different ways. In this text, we will only use the conformal bootstrap approach, and omit other historically relevant techniques such as Coulomb gas integrals, and the modular bootstrap. We believe that the conformal bootstrap approach is technically simpler, more general, and more powerful. In particular, our main objects are correlation functions on the sphere, and we do not have a rigid notion of which fields belong to a CFT. For example, in the Ising model, we can introduce disorder fields or compute cluster connectivities, without having to worry about modular invariance of the torus partition function.
Solving a CFT means determining its spectrum and structure constants. We deduce spectrums from assumptions such as
\begin{itemize}
\item the existence of degenerate fields,
\item singe-valuedness of correlation functions,
\item analyticity in the central charge or in conformal dimensions, which in particular allows us to find CFTs as limits of sequences of other CFTs.
\end{itemize}
These assumptions also lead to constraints on structure constants, which may or may not be enough for determining them analytically. If they are not enough, we can determine structure constants numerically by solving crossing symmetry equations. And from sufficiently precise numerical results, we can deduce analytic formulas.
\subsubsection*{This text's approach}
We strive to give simple derivations of non-trivial results, without introducing too many auxiliary objects that are later discarded. The calculations are not all done in detail in this text, but they are doable with reasonable effort. Readers who want formal exercises are directed to earlier texts in the same spirit: the minimal review \cite{rib16}, and the more extended review \cite{rib14}. However, these texts do not cover loop CFTs.
Technical terms that are defined in this text can be found in the index. We also use standard terms that we do not define, because adequate definitions are easily found in Wikipedia or other resources. For example, the definitions of an \textit{irreducible representation}, an \textit{indecomposable representation} or a \textit{projective representation} are found in Wikipedia as the first results of searches for these terms.
\subsubsection*{Acknowledgements}
\begin{itemize}
\item This text is based on lecture notes for a course given at IPhT Saclay in March-April 2024. Video recordings are available on \href{https://www.youtube.com/playlist?list=PLrLctLPAdPNtD7yFOcfdNx2J-A_CE1svr}{IPhT-TV}. I am grateful to the course organizers Riccardo Guida, Pierfrancesco Urbani and Monica Guica.
\item I would like to thank Linnea Grans-Samuelsson, Jesper Jacobsen, Santiago Migliaccio, Nikita Nemkov, Rongvoram Nivesvivat, Marco Picco, Paul Roux, Hubert Saleur, and Raoul Santachiara for collaboration on loop CFTs and related subjects.
\item I am grateful to Victor Godet, Jesper Jacobsen, Paul Roux, Ingo Runkel, Hubert Saleur, Xi Yin and Bernardo Zan for helpful discussions and correspondence.
\item I would like to thank Max Downing, Kay Wiese and Rongvoram Nivesvivat for comments and suggestions on the manuscript.
\end{itemize}
%\end{tcolorbox}
\section{Basics of 2d CFT}\label{sec:bo}
In this section we introduce conformal symmetry, fields, operator product expansions, and correlation functions. In particular, we emphasize the 2 ingredients of exact solvability: \textit{local} conformal symmetry and \textit{degenerate} fields. We begin with the Virasoro algebra, which describes local conformal transformations.
\subsection{The Virasoro algebra and its representations}\label{sec:vir}
\subsubsection{The Virasoro algebra}
Let us consider the Riemann sphere $\overline{\mathbb{C}}=\mathbb{C}\cup \{\infty\}$, equipped with the metric $ds^2 = dzd\bar z$. By definition, a \myindex{conformal transformation}\index{conformal transformation} of a Riemannian manifold is a transformation that preserves angles. On any open subset of the Riemann sphere, any holomorphic map $z\to f(z)$ is conformal, because it transforms the metric into itself, up to a scalar factor:
\begin{align}
dzd\bar z\to dfd\bar f = |f'(z)|^2 dzd\bar z\ .
\end{align}
Conversely, any conformal map is holomorphic.
In a quantum field theory on the Riemann sphere, states live on a constant time slice $|z|=1$, so we consider conformal transformations of that circle, equivalently of $\mathbb{C}^*= \mathbb{C}\backslash \{0\}$.
Let the \myindex{Witt algebra}\index{Witt algebra} be the algebra of infinitesimal conformal transformation of $\mathbb{C}^*$: this algebra is infinite-dimensional, with the basis
\begin{align}
\left(\ell_n\right)_{n\in\mathbb{Z}} \quad \text{with} \quad \ell_n = -z^{n+1}\frac{\partial}{\partial z}\ ,
\label{lpz}
\end{align}
and the commutation relations
\begin{align}
[\ell_n,\ell_m] = (n-m)\ell_{m+n}\ .
\end{align}
The generators of the Witt algebra include the translation generator $\ell_{-1} = -\frac{\partial}{\partial z}$, and the dilation generator $\ell_0 = -z\frac{\partial}{\partial z}$. In fact, $(\ell_{-1},\ell_0,\ell_1)$ generate the infinitesimal conformal transformations of the Riemann sphere.
The corresponding Lie group is the group of \myindex{global conformal transformations}\index{conformal transformation!---global} $PSL_2(\mathbb{C})$, whose elements act as
\begin{align}
z \mapsto \frac{az+b}{cz+d}\quad , \quad (a,b,c,d\in \mathbb{C},\ ad-bc\neq 0)\ .
\label{abcd}
\end{align}
Now, in a quantum theory, symmetries act projectively on states. And projective representations of an algebra are equivalent to representations of that algebra's central extension.
Therefore, the algebra that describes local conformal transformations in conformal field theory is the Witt algebra's central extension, called the Virasoro algebra. The \myindex{Virasoro algebra}\index{Virasoro algebra} $\mathfrak{V}$ has generators $(L_n)_{n\in\mathbb{Z}}$ that correspond to the Witt algebra generators, plus a central generator $\mathbf 1$. The commutation relations are
\begin{align}
[\mathbf 1, L_n] = 0 \quad , \quad \boxed{[L_n,L_m] = (n-m)L_{n+m} +\frac{c}{12}(n-1)n(n+1)\delta_{n+m,0}\mathbf 1} \ .
\label{vir}
\end{align}
The \myindex{central charge}\index{central charge} $c\in\mathbb{C}$ is a fundamental parameter not only of the Virasoro algebra, but also of any conformal field theory built on that algebra.
\subsubsection{Highest-weight representations}\label{sec:hwr}
In a conformal field theory, the space of states is a representation of the Virasoro algebra, and can be decomposed into indecomposable representations. But which indecomposable representations are physically relevant? To answer this question, we will focus on the properties of the dilation generator $L_0$ of the Virasoro algebra. Conceptually, this is because $L_0$ can be interpreted as the energy operator. Technically, this is because $L_0$ controls the convergence of operator product expansions, as we will see in Section \ref{sec:csope}. More specifically, a necessary condition for convergence is that the eigenvalues of $L_0$ be bounded from below in any indecomposable representation. We will always assume that this condition holds.
The eigenvalues of $L_0$ are called \myindex{conformal dimensions}\index{conformal!---dimension}. Under the action of the Virasoro generator $L_n$, the conformal dimension decreases by $n$. For any vector $V$ in a representation of the Virasoro algebra, we indeed have
\begin{align}
L_0V{} = \Delta V{} \quad \Rightarrow\quad L_0 L_nV{} = L_nL_0V{} + [L_0, L_n] V{} = (\Delta-n)L_nV{} \ .
\end{align}
In an indecomposable representation $\mathcal{R}$, the $L_0$-eigenstate with the lowest eigenvalue must therefore be annihilated by $L_{n>0}$. This eigenstate is therefore a \myindex{primary state}\index{primary!---state}, where we define a primary state $V_\Delta$ of conformal dimension $\Delta$ by
\begin{align}
\boxed{L_0 V_\Delta = \Delta V_\Delta \quad , \quad L_{n>0} V_\Delta = 0}\ .
\end{align}
Let us introduce a basis of \myindex{creation operators}\index{creation operator},
\begin{align}
\mathcal{L} = \left\{\prod_{i=1}^k L_{-n_i} \right\}_{k\in\mathbb{N},\ 0<n_1\leq \dots \leq n_k}=\left\{1, L_{-1}, L_{-1}^2, L_{-2},\cdots \right\}\ .
\label{lcm}
\end{align}
Any state of the type $L V{}$ with $L\in\mathcal{L}$ is called a \myindex{descendant state}\index{descendant!---state} of the primary state $V$. Let $N=|L|=\sum_{i=1}^k n_i \in\mathbb{N}$ be the \myindex{level}\index{level} of that state, then its conformal dimension is $\Delta+N$, and we write
\begin{align}
\mathcal{L}_N = \Big\{ L\in\mathcal{L}\Big| |L|=N\Big\}\ .
\label{ln}
\end{align}
By extension, a linear combination of descendant states is also called a descendant state.
A primary state $V{} \in \mathcal{R}$ generates a subrepresentation $\mathcal{R}_{V{}}= \text{Span}(LV{})_{L\in\mathcal{L}}\subset\mathcal{R}$, which is the space of its descendant states. If the states $(LV{})_{L\in\mathcal{L}}$ are linearly independent, then $\mathcal{R}_{V{}}$ is called the \myindex{Verma module}\index{Verma module} $\mathcal{V}_\Delta$. If not, $\mathcal{R}_{V{}}$ is a quotient of $\mathcal{V}_\Delta$ by a subrepresentation: such quotients are called \myindex{highest-weight representations}\index{highest-weight representation}. In such a representation, $L_0$ is diagonalizable.
Let us sketch a Verma module by displaying all its basis states up to the level $N=4$, with arrows representing the action of the Virasoro generators $L_{-1},L_{-2},L_{-3},L_{-4}$:
\begin{align}
\begin{tikzpicture}[scale = .3, baseline=(current bounding box.center)]
\draw[-latex, very thick] (20, 0) -- (20, -27) node [right] {$N$};
\foreach \x in {0, ..., 4}
{
\draw [dotted] (-20, {-6*\x}) -- (20, {-6*\x}) node [right] {${\x}$};
}
\node[fill = white] at (0, 0) (0) {$1$};
\node[fill = white] at (-3,-6) (1) {$L_{-1}$};
\node[fill = white] at (-6, -12) (11) {$L_{-1}^2$};
\node[fill = white] at (-9, -18) (111) {$L_{-1}^3$};
\node[fill = white] at (-12, -24) (1111) {$L_{-1}^4$};
\node[fill = white] at (0,-12) (2) {$L_{-2}$};
\node[fill = white] at (-3,-18) (12) {$L_{-1}L_{-2}$};
\node[fill = white] at (-6,-24) (112) {$L_{-1}^2L_{-2}$};
\node[fill = white] at (0,-24) (22) {$L_{-2}^2$};
\node[fill = white] at (6,-18) (3) {$L_{-3}$};
\node[fill = white] at (6,-24) (13) {$L_{-1}L_{-3}$};
\node[fill = white] at (12,-24) (4) {$L_{-4}$};
\draw[-latex] (0) -- (1);
\draw[-latex] (1) -- (11);
\draw[-latex] (11) -- (111);
\draw[-latex] (111) -- (1111);
\draw[-latex] (0) -- (2);
\draw[-latex] (0) -- (3);
\draw[-latex] (0) -- (4);
\draw[-latex] (2) -- (12);
\draw[-latex] (12) -- (112);
\draw[-latex] (2) -- (22);
\draw[-latex] (3) -- (13);
\end{tikzpicture}
\label{verma}
\end{align}
States such as
$L_{-2}L_{-1}V = L_{-1}L_{-2}V -L_{-3}V$ are linear combination of elements of our basis, and are therefore not displayed.
\subsubsection{Null vectors of Verma modules}\label{sec:nvvm}
In a representation of the Virasoro algebra, we define a \myindex{null vector}\index{null vector} or singular vector as a primary state that is a descendant of another primary state. Null vectors play an important structural role: a highest-weight representation is irreducible if and only if it has no null vector.
Let us look for null vectors in the Verma module $\mathcal{V}_\Delta$. This can be done level by level, starting with level $1$. That level is generated by a single state $L_{-1}V_\Delta$. To determine whether that state is primary, let us determine how it behaves under the action of $L_{n\geq 1}$:
\begin{align}
L_n L_{-1}V_\Delta = [L_n, L_{-1}] V_\Delta = (n+1) L_{n-1}V_\Delta =
\left\{\begin{array}{ll} 0 & \quad \text{if } n\geq 2\ , \\ 2\Delta V_\Delta & \quad \text{if } n = 1\ . \end{array}\right.
\end{align}
Therefore, $\mathcal{V}_\Delta$ has a null vector at level $1$ if and only if $\Delta=0$. Next, consider the case of level $2$, where we write states as
\begin{align}
X{} = \left(a^{L_{-1}^2} L_{-1}^2 + a^{L_{-2}} L_{-2}\right) V_\Delta\ ,
\end{align}
where $a^{L_{-1}^2}$ and $a^{L_{-2}}$ are complex coefficients. It is easy to show that $L_{n\geq 3}X{}=0$, and we compute
\begin{subequations}
\label{l1x}
\begin{align}
L_1X{} &= \left((4\Delta+2)a^{L_{-1}^2} + 3a^{L_{-2}}\right) L_{-1}V_\Delta\ ,
\\
L_2 X{} &= \left(6\Delta a^{L_{-1}^2}+(4\Delta+\tfrac12 c) a^{L_{-2}}\right)V_\Delta\ .
\end{align}
\end{subequations}
Therefore, the conditions $L_1X{}=L_2 X{}=0$ for $X{}$ to be a null vector boil down to a system of 2 linear equations for the coefficients $(a^{L_{-1}^2},a^{L_{-2}})$. The vanishing of that system's determinant determines the conformal dimension $\Delta$ as a function of the central charge $c$,
\begin{align}
\Delta = \frac{1}{16}\left( 5-c\pm\sqrt{(c-1)(c-25)} \right) \ .
\label{dpm}
\end{align}
Solving CFTs will lead to fairly complicated formulas, which would become even more complicated if we tolerated a square root at this early stage.
To get rid of the square root, we rewrite the central charge in terms of the parameter $\beta$ such that
\begin{align}
\boxed{c = 1- 6\left(\beta - \beta^{-1}\right)^2 } \ .
\label{cb}
\end{align}
Then the conformal dimensions \eqref{dpm} become
$
\Delta = -\frac12 + \frac34\beta^{\pm 2}
$, and the corresponding null vectors are $X\propto \left(L_{-1}^2 -\beta^{\pm 2}L_{-2}\right)V_\Delta$.
The price to pay for this simplification is that the 4 values $\beta,-\beta,\beta^{-1},-\beta^{-1}$ all correspond to the same central charge.
More generally, there is an infinite family of conformal dimensions $(\Delta_{(r, s)})_{r,s\in\mathbb{N}^*}$ such that $\mathcal{V}_{\Delta_{(r, s)}}$ has a null vector $L_{\langle r,s\rangle} V_{\Delta_{(r, s)}}$ at level $N=rs$. Let us summarize the cases $N=1,2,3,4$ in these notations:
\begin{align}
\renewcommand{\arraystretch}{1.5}
\begin{array}{|c|c|c|l|}
\hline
N & (r,s) & \Delta_{(r,s)} & L_{\langle r,s\rangle}
\\
\hline\hline
1 & (1,1) & 0 & L_{-1}
\\
\hline
\multirow{2}{*}{2} &
(2,1) & -\frac12 +\frac{3}{4} \beta^2 & L_{-1}^2 -\beta^2 L_{-2}
\\
\cline{2-4}
& (1,2) & -\frac12 + \frac{3}{4}\beta^{-2} & L_{-1}^2 -\beta^{-2} L_{-2}
\\
\hline
\multirow{2}{*}{3} &
(3,1 ) & -1 +2 \beta^2 &L_{-1}^3 -4\beta^2 L_{-1}L_{-2}+2\beta^2(\beta^2+1)L_{-3}
\\
\cline{2-4}
& (1,3 ) & -1 +2 \beta^{-2} &L_{-1}^3 -4\beta^{-2} L_{-1}L_{-2}+2\beta^{-2}(\beta^{-2}+1)L_{-3}
\\
\hline
\multirow{5}{*}{4} &
(4,1 ) & -\frac32 +\frac{15}{4} \beta^2 &
\begin{array}{r}
L_{-1}^4 -10\beta^2L_{-1}^2L_{-2}
+2\beta^2\left(12\beta^2+5\right) L_{-1}L_{-3}
\\
+9\beta^4 L_{-2}^2
-6\beta^2\left(6\beta^4+4\beta^2+1\right)L_{-4}
\end{array}
\\
\cline{2-4}
& (2,2) & \frac34\left(\beta-\beta^{-1}\right)^2 &
\begin{array}{l}
L_{-1}^4-2\left(\beta^2+\beta^{-2}\right)L_{-1}^2L_{-2}
+\left(\beta^2-\beta^{-2}\right)^2 L_{-2}^2
\\
+2\left(1+\left(\beta+\beta^{-1}\right)^2\right) L_{-1}L_{-3}
-2\left(\beta+\beta^{-1}\right)^2 L_{-4}
\end{array}
\\
\cline{2-4}
& (1,4 ) & -\frac32 +\frac{15}{4} \beta^{-2} &
\begin{array}{r}
L_{-1}^4 -10\beta^{-2}L_{-1}^2L_{-2}
+2\beta^{-2}\left(12\beta^{-2}+5\right) L_{-1}L_{-3}
\\
+9\beta^{-4} L_{-2}^2
-6\beta^{-2}\left(6\beta^{-4}+4\beta^{-2}+1\right)L_{-4}
\end{array}
\\
\hline
\end{array}
\label{ars}
\end{align}
To write $\Delta_{(r, s)}$ for any $r,s\in\mathbb{N}^*$, we rewrite the conformal dimension in terms of the momentum $P$ defined by
\begin{align}
\boxed{\Delta = \frac{c-1}{24} + P^2}\ .
\label{dp}
\end{align}
We can rewrite Verma modules as $\mathcal{V}_P=\mathcal{V}_{-P}$, where the parameter $P$ is redundant due to the invariance of $\Delta$ under the \myindex{reflection}\index{reflection} $P\to -P$. The dimensions $\Delta_{(r, s)}$ correspond to the momentums
\begin{align}
\boxed{P_{(r, s)} = \frac12\left(r\beta -s\beta^{-1}\right)}\ ,
\label{prs}
\end{align}
and it follows that the null vector $L_{\langle r,s\rangle} V_{\Delta_{(r, s)}}$ is a primary state of dimension
\begin{align}
\boxed{\Delta_{(r, s)} + rs = \Delta_{(r, -s)}}\ ,
\label{drms}
\end{align}
We will provide a derivation of $P_{(r, s)}$ using operator product expansions in Section \ref{sec:dope}. On the other hand, the expressions of $L_{\langle r,s\rangle} $ for $rs\geq 3$ are complicated and not particularly useful, see \cite{ken92} or \cite[Appendix 8A]{fms97}.
For $\beta^2\in\mathbb{C}\backslash \mathbb{Q}$, the Verma module $\mathcal{V}_\Delta$ has a null vector if and only if $\Delta=\Delta_{(r, s)}$ for some $r,s\in\mathbb{N}^*$, in which case it has only 1 null vector. For $\beta^2\in\mathbb{Q}$ however, a Verma module can have several null vectors. If $\beta^2 = \frac{q}{p}$ with $p,q$ coprime integers, we indeed have the identity
\begin{align}
\Delta_{(r+p,s+q)} = \Delta_{(r,s)}\ ,
\label{rpsq}
\end{align}
which together with $\Delta_{(r,s)}=\Delta_{(-r,-s)}$ implies coincidences of the type $\Delta_{(r,s)}=\Delta_{(r',s')}$ with $r',s'\in\mathbb{Z}$.
Given $r_1,s_1\in \mathbb{N}^*$, the reducible Verma module $\mathcal{V}_{\Delta_{(r_1,s_1)}}$ has a null vector $L_{\langle r_1,s_1\rangle} V_{\Delta_{(r,s)}}$. If $pq>0$, or if $pq<0$ with $\left\lfloor \frac{r_1}{|p|}\right\rfloor \neq \left\lfloor \frac{s_1}{|q|}\right\rfloor$ and $\left\lceil \frac{r_1}{|p|}\right\rceil \neq \left\lceil \frac{s_1}{|q|}\right\rceil$, then the dimension $\Delta_{(r_1,-s_1)}$ \eqref{drms} of this null vector is of the type $\Delta_{(r_2,s_2\rangle}$ with $r_2,s_2\in\mathbb{N}^*$, leading to another null vector $L_{\langle r_2,s_2\rangle}L_{\langle r_1,s_1\rangle} V_{\Delta_{(r,s)}}$.
It turns out that all null vectors of Verma modules are of the type $L_{\langle r_1,s_1\rangle} V_{\Delta_{(r,s)}}$ or $L_{\langle r_2,s_2\rangle}L_{\langle r_1,s_1\rangle} V_{\Delta_{(r,s)}}$. In particular, given a null vector of the type $L_{\langle r_3,s_3\rangle}L_{\langle r_2,s_2\rangle}L_{\langle r_1,s_1\rangle} V_{\Delta_{(r_1,s_1)}}$,
\begin{align}
\exists r_4,s_4\in \mathbb{N}^*\ , \ L_{\langle r_3,s_3\rangle}L_{\langle r_2,s_2\rangle}L_{\langle r_1,s_1\rangle} V_{\Delta_{(r_1,s_1)}} = L_{\langle r_4,s_4\rangle} V_{\Delta_{(r_4,s_4)}}\ .
\label{lllel}
\end{align}
To see this, let $\epsilon,\eta\in\{+,-\}$ be the signs such that $P_{(-r_1,s_1)}=\epsilon P_{(r_2,s_2)}$ and $P_{(r_2,-s_2)}=\eta P_{(r_3,s_3)}$, and let $r_4=|\epsilon r_1-r_2+\eta r_3|$ and $s_4=|\epsilon s_1+s_2+\eta s_3|$, then $\sum_{i=1}^3 r_is_i=r_4s_4$, i.e. the 2 null vectors in Eq. \eqref{lllel} have the same level.
To summarize, numbers of null vectors depend a lot on $\beta^2$. If $\beta^2<0$, then $\{\Delta_{(r, s)}\}_{r,s\in\mathbb{N}^*}$ is bounded from above, and Eq. \eqref{drms} implies that a Verma module can only have finitely many null vectors. If $\beta^2\in \mathbb{Q}_{>0}$, then by Eq. \eqref{rpsq} the existence of a null vector implies the existence of infinitely many others:
\begin{align}
\begin{array}{|c||c|c|c|}
\hline
\text{Value of }\beta^2 & \mathbb{C}\backslash \mathbb{Q} & \mathbb{Q}_{>0} & \mathbb{Q}_{<0}
\\
\hline
\text{Value of }c & \text{generic} & c\leq 1 , c\in\mathbb{Q} & c\geq 25, c\in\mathbb{Q}
\\
\hline
\#\text{null vectors in } \mathcal{V}_{\Delta_{(r, s)}} & 1 & \infty & \text{finite}
\\
\hline
\end{array}
\end{align}
To illustrate the 2 cases $\beta^2\in \mathbb{Q}_{>0}$ and $\beta^2\in \mathbb{Q}_{<0}$, let us plot at each $(r,s)\in \{1,\dots, 10\}^2$ a dot whose size increases with
$\Delta_{(r,s)}$, together with the vector $(p, q)$ of Eq. \eqref{rpsq} as a red arrow:
\begin{align}
\begin{tikzpicture}[baseline=(base), scale = .5]
\coordinate (base) at (0, 4);
\draw[latex-latex] (0, 10) node [left] {$s$} -- (0, 0) node[below left] {$0$} -- (10, 0) node [below] {$r$};
\draw[thick, red, -latex] (0, 0) -- (5, 3);
\foreach \r in {1,...,10}{
\foreach \s in {1,...,10}{
%\pgfmathsetmacro{\mycol}{100-1.8*abs(3*\r-5*\s)}
%\filldraw[black!\mycol] (\r, \s) circle [radius = 4pt];
\pgfmathsetmacro{\myrad}{.2+.12*abs(3*\r-5*\s)}
\filldraw (\r, \s) circle [radius = \myrad pt];
}}
\node at (5, -2) {$\beta^2 = \frac35>0$};
\end{tikzpicture}
\qquad \qquad
\begin{tikzpicture}[baseline=(base), scale = .5]
\coordinate (base) at (0, 4);
\draw[latex-latex] (0, 10) node [left] {$s$} -- (0, 0) node[below left] {$0$} -- (10, 0) node [below] {$r$};
\draw[thick, red, -latex] (5, 0) -- (0, 3);
\foreach \r in {1,...,10}{
\foreach \s in {1,...,10}{
%\pgfmathsetmacro{\mycol}{3*\r+5*\s}
%\filldraw[black!\mycol] (\r, \s) circle [radius = 4pt];
\pgfmathsetmacro{\myrad}{6-(3*\r+5*\s)/15}
\filldraw (\r, \s) circle [radius = \myrad pt];
}}
\node at (5, -2) {$\beta^2 = -\frac35<0$};
\end{tikzpicture}
\end{align}
\subsubsection{Vanishing and non-vanishing null vectors}
When a representation $\mathcal{R}$ has a null vector $X{}=LV{}$, this allows us to define a quotient representation $\frac{\mathcal{R}}{\mathcal{R}_{X{}}}$ by setting the null vector to zero.
By an abuse of terminology, we say that $\frac{\mathcal{R}}{\mathcal{R}_{X{}}}$ has a \myindex{vanishing null vector}\index{null vector!vanishing---}, which really means that its primary state $V{}$ obeys the \myindex{null vector equation}\index{null vector!---equation} $LV{}=0$. On the other hand, $\mathcal{R}$ itself has a non-vanishing null vector $X{}$.
If $\mathcal{R}$ has several null vectors, we can define various quotient representations by setting some null vectors to zero. A \myindex{degenerate representation}\index{degenerate!---representation} is a representation that has at least 1 vanishing null vector. A degenerate representation is partly degenerate if it has some non-vanishing null vectors, and \myindex{fully degenerate}\index{fully degenerate!---representation} if all null vectors vanish. Partly degenerate representations play an important role in CFTs with extended chiral algebras, such as conformal Toda theories \cite{fl07c}.
For $r,s\in\mathbb{N}^*$, the level-$rs$ null vector $L_{\langle r,s\rangle}V_{\Delta_{(r,s)}}\in \mathcal{V}_{\Delta_{(r,s)}}$ is a primary state of dimension $\Delta_{(r,-s)}$, according to Eq. \eqref{drms}. Together with its descendants, this state generates a Verma module $\mathcal{V}_{\Delta_{(r,-s)}}\subset \mathcal{V}_{\Delta_{(r,s)}}$. Setting the null vector to zero, we obtain the degenerate representation
\begin{align}
\mathcal{R}^d_{\langle r,s\rangle}
=\frac{\mathcal{V}_{\Delta_{(r,s)}}}{\mathcal{V}_{\Delta_{(r,-s)}}}\ .
\label{rvv}
\end{align}
If $\beta^2\in \mathbb{Q}$, the Verma module $\mathcal{V}_{\Delta_{(r,s)}}$ in general has several null vectors, and $\mathcal{R}^d_{\langle r,s\rangle}$ may be partly degenerate. Let us call $\mathcal{R}^f_{\langle r,s\rangle}$ the fully degenerate quotient of $\mathcal{V}_{\Delta_{(r,s)}}$: this quotient can generally not be obtained by setting 1 null vector to zero, but it is always enough to set 2 null vectors to zero, with all other null vectors being descendants of these two. Schematically,
\begin{align}
\mathcal{R}^f_{\langle r,s\rangle} = \frac{\mathcal{V}_{\Delta_{(r,s)}}}{\mathcal{V}_{\Delta_{(r_1,-s_1)}}+ \mathcal{V}_{\Delta_{(r_2,-s_2)}}}\ .
\end{align}
We define the \myindex{degenerate state}\index{degenerate!---state} $V^d_{\langle r,s\rangle}$ and \myindex{fully degenerate state}\index{fully degenerate!---state} $V^f_{\langle r,s\rangle}=V^f_{\langle r_1,s_1\rangle}=V^f_{\langle r_2,s_2\rangle}$ as the primary states of the representation $\mathcal{R}^d_{\langle r,s\rangle}$ and
$\mathcal{R}^f_{\langle r,s\rangle}$. These states obey the null vector equation
\begin{align}
L_{\langle r, s\rangle} V^d_{\langle r,s\rangle} = L_{\langle r, s\rangle} V^f_{\langle r,s\rangle} = 0\ .
\label{lvdz}
\end{align}
We use the notation $V_{\Delta_{(r,s)}}$ for a primary state that has the same dimension as $V^d_{\langle r,s\rangle}$ and $V^f_{\langle r,s\rangle}$, but does not necessarily obey the null vector equation.
A basis of the degenerate representation $\mathcal{R}^d_{\langle r,s\rangle}$ is formed by the descendant states $LV^d_{\langle r,s\rangle}$ such that the creation operator $L$ does not involve the Virasoro generator $L_{-rs}$. In the simplest example $(r,s)=(1,1)$, let us sketch $\mathcal{R}^d_{\langle 1,1\rangle}$ up to the level $N=4$, so that it may be compared to a Verma module \eqref{verma}:
\begin{align}
\begin{tikzpicture}[scale = .15, baseline=(current bounding box.center)]
\draw[-latex, very thick] (20, 0) -- (20, -28) node [right] {$N$};
\foreach \x in {0, ..., 4}
{
\draw [dotted] (-6, {-6*\x}) -- (20, {-6*\x}) node [right] {${\x}$};
}
\node[fill = white] at (0, 0) (0) {$1$};
\node[fill = white] at (0,-12) (2) {$L_{-2}$};
\node[fill = white] at (0,-24) (22) {$L_{-2}^2$};
\node[fill = white] at (6,-18) (3) {$L_{-3}$};
\node[fill = white] at (12,-24) (4) {$L_{-4}$};
\draw[-latex] (0) -- (2);
\draw[-latex] (0) -- (3);
\draw[-latex] (0) -- (4);
\draw[-latex] (2) -- (22);
\end{tikzpicture}
\label{r11}
\end{align}
\subsubsection{The Shapovalov form}
We will now introduce the Shapovalov form, which will be useful for computing operator product expansions and conformal blocks. We will see that the Shapovalov from has zeros related to null vectors, and this will lead to poles of the operator product expansions and conformal blocks.
Let us define an involution on the Virasoro algebra by
\begin{align}
L^*_n = L_{-n} \ .
\end{align}
This involution is a Lie algebra automorphism, as it obeys $[L_m,L_n]^* = -[L_m^*,L_n^*]$. It can be extended to the universal enveloping algebra by $\left(LL'\right)^* = L'^*L^*$. Using this involution, we define the \myindex{Shapovalov form}\index{Shapovalov form} on any highest-weight representation as the symmetric bilinear form such that
\begin{align}
\big(LV_\Delta\big| L'V_\Delta\big) = \left\{\begin{array}{ll} 0 & \text{if } |L|\neq |L'|\ ,
\\
S_{L,L'}(\Delta) & \text{if } |L|=|L'|\ , \quad \text{where} \quad L^*L'V_\Delta = S_{L,L'}(\Delta) V_\Delta\ .
\end{array}\right.
\label{shap}
\end{align}
At each level $N\in\mathbb{N}$, the Shapovalov form is described by a matrix that depends on $\Delta$ and $c$. In the cases $N=0,1,2$, these matrices are
\begin{align}
S_{1,1} = 1 \ \ , \ \ S_{L_{-1},L_{-1}} = 2\Delta \ \ , \ \ \begin{bmatrix} S_{L_{-1}^2,L_{-1}^2} & S_{L_{-1}^2,L_{-2}} \\ S_{L_{-1}^2,L_{-2}} & S_{L_{-2},L_{-2}}\end{bmatrix} =
\begin{bmatrix} 4\Delta(2\Delta+1) & 6\Delta \\ 6\Delta & 4\Delta+\frac{c}{2}\end{bmatrix}\ .
\label{sh012}
\end{align}
From its definition, it follows that the Shapovalov form has zeros:
\begin{align}
\forall L, L'\in \mathcal{L}\ , \quad S_{LL_{\langle r,s\rangle},L'}(\Delta_{(r,s)})=0 \ .
\end{align}
For generic values of the central charge, these are simple zeros. This implies that the inverse Shapovalov matrix at level $N$ has simple poles for $\Delta\in \left\{\Delta_{(r,s)}\right\}_{rs\leq N}$. In a basis of level-$rs$ creation operators that includes $L_{\langle r,s\rangle}$, the level-$rs$ inverse Shapovalov matrix behaves as
\begin{align}
S^{-1}_{L,L'}(\Delta) \underset{\Delta\to \Delta_{(r,s)}} = \frac{1}{\Delta-\Delta_{(r,s)}} \frac{\delta_{L,L_{\langle r,s\rangle}}\delta_{L',L_{\langle r,s\rangle}}}{S'_{L_{\langle r,s\rangle},L_{\langle r,s\rangle}}(\Delta_{(r,s)})} + O(1) \ .
\label{smo}
\end{align}
In fact, the level-$N$ inverse Shapovalov matrix can be computed explicitly in terms of $\left\{L_{\langle r,s\rangle}\right\}_{rs\leq N}$ \cite{fqs24}.
Using the same coefficients \eqref{shap}, we can define a sesquilinear form instead of a bilinear form. This sesquilinear form is Hermitian if its coefficients are real, i.e. if $\Delta,c\in\mathbb{R}$. Our highest-weight representation is called \myindex{unitary}\index{highest-weight representation!unitary---} if that Hermitian form is positive definite. Unitarity plays an important role in some applications of CFT, but it has no bearing on exact solvability, so we will not elaborate further on that subject.
\subsection{Fields and operator product expansions}\label{sec:fope}
The fundamental objects of conformal field theory are fields. In quantum field theory, all states can be obtained by acting on the vacuum state with fields: this is called the \myindex{state-field correspondence}\index{state-field correspondence} or operator-state correspondence. In conformal field theory, the correspondence is bijective, and allows us to identify the space of fields with the \myindex{space of states}\index{space of states} or \myindex{spectrum}\index{spectrum} --- a vector space on which the Virasoro algebra acts. To a state $V$, the correspondence associates a field $V(z)$, which depends on a position $z\in\overline{\mathbb{C}}$ in the Riemann sphere. Primary states, descendant states and degenerate states respectively give rise to \myindex{primary fields}\index{primary!---field}\index{field!primary---}, \myindex{descendant fields}\index{descendant!---field}\index{field!descendant---} and \myindex{degenerate fields}\index{degenerate!---field}\index{field!degenerate---}.
In the bootstrap approach, which is axiomatic rather than constructive, we need not define fields. Rather, we use fields as a convenient notation for stating properties of correlation functions, which we will introduce in Section \ref{sec:cor}.
\subsubsection{The energy-momentum tensor}
The dependence of fields on the coordinate $z$ is determined by the assumption that the Virasoro generator $L_{-1}$ generates translations, just like the Witt algebra generator $\ell_{-1}$ \eqref{lpz}:
\begin{align}
\boxed{\frac{\partial}{\partial z} V(z) = L_{-1} V(z)} \ .
\label{pvlv}
\end{align}
Since fields depend on $z$, the Virasoro algebra that acts on fields also depends on $z$. This can be made explicit using the notation
\begin{align}
L_n V(z) = L_n^{(z)} V(z) \ .
\end{align}
Applying Eq. \eqref{pvlv} to $L_n^{(z)} V(z)$, we find how the Virasoro generator $L_n^{(z)}$ depends on $z$,
\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}
This can be rewritten as the $z$-independence of a formal Laurent series that combines all the generators $(L_n^{(z)})_{n\in\mathbb{Z}}$,
\begin{align}
\frac{\partial}{\partial z} \sum_{n\in\mathbb{Z}} \frac{L_n^{(z)}}{(y-z)^{n+2}} = 0\ ,\qquad (\forall y\in\mathbb{C}\backslash \{z\})\ .
\end{align}
The formal Laurent series is then called the \myindex{energy-momentum tensor}\index{energy-momentum tensor},
\begin{align}
\boxed{T(y) = \sum_{n\in\mathbb{Z}} \frac{L_n^{(z)}}{(y-z)^{n+2}}} \ .
\label{tl}
\end{align}
A priori, the energy-momentum tensor only makes sense when acting on a field $V(z)$,
\begin{align}
T(y)V(z) = \sum_{n\in\mathbb{Z}} \frac{L_n V(z)}{(y-z)^{n+2}}\quad , \quad L_n V(z) = \frac{1}{2\pi i} \oint_{z}dy\ (y-z)^{n+1} T(y)V(z)\ .
\label{lvtv}
\end{align}
However, we will shortly interpret $T(y)$ as a field, and $T(y)V(z)$ as an operator product expansion of 2 fields. The statement that $V_\Delta(z)$ is a primary field of dimension $\Delta$ is equivalent to the OPE
\begin{align}
\boxed{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{tvd}
\end{align}
where we used Eq. \eqref{pvlv} and wrote the regular terms as $O(1)$.
\subsubsection{Operator product expansions}\label{sec:ope}
The fundamental axiom of conformal field theory, which underlies the conformal bootstrap approach, is the existence of \myindex{operator product expansions (OPEs)}\index{OPE}. This axiom states that there exists a bilinear product on the space of fields,
\begin{align}
\boxed{V_1(z_1)V_2(z_2) = \sum_{k\in\mathcal{S}} C^k_{12}(z_1,z_2) V_k(z_2)}\ ,
\label{ope}
\end{align}
for some family $(V_k)_{k\in\mathcal{S}}$ of linearly independent states, and
some functions $C^k_{12}(z_1,z_2)$ called \myindex{OPE coefficients}\index{OPE!---coefficient}. Operator product expansions generally exist in quantum field theory, but they are particularly useful in conformal field theory because they are supposed to hold for any $z_1$ in a finite neighbourhood of $z_2$ (but $z_1\neq z_2$), and not just asymptotically as $z_1\to z_2$. The finite neighbourhood in question may depend on the correlation function in which we perform the OPE, and also on the basis $(V_k)_{k\in\mathcal{S}}$.
We further assume that the product of fields is commutative,
\begin{align}
\boxed{V_1(z_1)V_2(z_2) = V_2(z_2)V_1(z_1)}\ .
\label{comm}
\end{align}
This implies that OPEs are \myindex{commutative}\index{OPE!---commutativity} and \myindex{associative}\index{OPE!---associativity}. However, the resulting constraints on the coefficients $C^i_{12}(z_1,z_2)$ are not particularly simple, because of having to choose a position ($z_2$ in our conventions) for the fields in the right-hand side of the OPE \eqref{ope}.
We also assume that there exists an \myindex{identity field}\index{identity field} $I(z)$ whose OPE with any other field is trivial, namely $I(z_1)V(z_2) = V(z_2)$. (The corresponding state is called the vacuum state.) Let us consider the energy-momentum tensor $T$ as a field, then we have
$
T(y)I(z) = T(y) = T(z) + O(y-z)
$.
Comparing this with Eq. \eqref{lvtv} applied to $I(z)$, we find
\begin{align}
L_{n\geq -1} I(z) = 0 \quad , \quad T(z) = L_{-2}I(z)\ .
\end{align}
In particular, the identity field is a degenerate primary field of conformal dimension zero, with a vanishing null vector at level $1$, i.e. $I(z) \propto V^d_{\langle 1,1\rangle}(z)$. And from the expression of $T(z)$ as a descendant of that field, we deduce
\begin{align}
T(y)T(z) = \frac{\frac{c}{2}}{(y-z)^4} + \frac{2T(z)}{(y-z)^2} + \frac{\partial T(z)}{y-z} + O(1)\ .
\label{tt}
\end{align}
This OPE is equivalent to the commutation relations of the Virasoro algebra \eqref{vir}.
\subsubsection{Conformal symmetry constraints on OPEs}\label{sec:csope}
For technical simplicity, we focus on an OPE of 2 primary fields. In that OPE, we single out the contributions of a primary field $V_\Delta$ and its descendants:
\begin{align}
V_{\Delta_1}(z_1)V_{\Delta_2}(z_2) \supset \sum_{L\in \mathcal{L}} C_{\Delta_1,\Delta_2}^{\Delta, L}(z_1,z_2) LV_\Delta(z_2) \ ,
\end{align}
where $\mathcal{L}$ \eqref{lcm} is the set of creation operators. On both sides of the OPE, let us insert $\frac{1}{2\pi i}\oint_{z_1,z_2} dy (y-z_2)^{n+1}T(y)$, where the integration contour encloses both $z_1$ and $z_2$. Using Eqs. \eqref{lvtv} and \eqref{tvd}, we find
\begin{align}
\left(L_n^{(z_2)} +z_{12}^{n+1}\partial_{z_1} + (n+1)z_{12}^n\Delta_1\right) V_{\Delta_1}(z_1)V_{\Delta_2}(z_2) \supset \sum_{L\in \mathcal{L}} C_{\Delta_1,\Delta_2}^{\Delta, L}(z_1,z_2)L_nLV_\Delta(z_2) \ .
\end{align}
We focus on 3 cases:
\begin{itemize}
\item $\boxed{n=-1}$\ : We use Eq. \eqref{pvlv} for $V_{\Delta_2}(z_2)$ and $LV_{\Delta}(z_2)$, and perform the OPE again on the left-hand side:
\begin{align}
\left(\partial_{z_1}+\partial_{z_2}\right) \sum_{L\in \mathcal{L}} C_{\Delta_1,\Delta_2}^{\Delta, L}(z_1,z_2)LV_\Delta(z_2) = \sum_{L\in \mathcal{L}} C_{\Delta_1,\Delta_2}^{\Delta, L}(z_1,z_2)\partial_{z_2} LV_\Delta(z_2) \ .
\end{align}
This leads to $(\partial_{z_1}+\partial_{z_2})C^{\Delta,L}_{\Delta_1,\Delta_2}(z_1,z_2)=0$, i.e. translation invariance of the OPE coefficients.
\item $\boxed{n=0}$\ : We use $L_0LV_{\Delta}(z_2)= (\Delta+|L|) V_{\Delta}(z_2)$, and perform the OPE again on the left-hand side. This leads to $(z_{12}\partial_{z_1}+\Delta_1+\Delta_2-\Delta-|L|)C^{\Delta,L}_{\Delta_1,\Delta_2}(z_1,z_2) = 0$, and we deduce the dependence of OPE coefficients on $z_1,z_2$:
\begin{align}
C^{\Delta,L}_{\Delta_1,\Delta_2}(z_1,z_2) = z_{12}^{\Delta+|L|-\Delta_1-\Delta_2} C^{\Delta,L}_{\Delta_1,\Delta_2}\ .
\end{align}
\item $\boxed{n\geq 1}$\ : We use $L_n^{(z_2)}V_{\Delta_2}(z_2)=0$, and perform the OPE again on the left-hand side. The action of the differential operator on the OPE coefficients is now known, and we find
\begin{align}
\sum_{L\in\mathcal{L}} C^{\Delta,L}_{\Delta_1,\Delta_2}z_{12}^{|L|+n}(\Delta+|L|+n\Delta_1-\Delta_2) LV_{\Delta}= \sum_{L\in\mathcal{L}} C^{\Delta,L}_{\Delta_1,\Delta_2}z_{12}^{|L|}L_n LV_{\Delta}\ .
\end{align}
Extracting the coefficient of $z_{12}^N$ for some $N\in\mathbb{N}$, we find
\begin{align}
L_n W_N = \theta_{N,n}W_{N-n} \quad \text{with} \quad \left\{\begin{array}{l} W_N = \sum_{|L|=N}C^{\Delta,L}_{\Delta_1,\Delta_2} LV_{\Delta}\ , \\ \theta_{N,n} = N-n+n\Delta_1-\Delta_2 + \Delta\ . \end{array}\right.
\label{opew}
\end{align}
\end{itemize}
These linear equations for the OPE coefficients $C^{\Delta,L}_{\Delta_1,\Delta_2}$ are called \myindex{OPE Ward identities}\index{Ward identity!OPE---}.
Let us look for a solution $\left(f^{\Delta,L}_{\Delta_1,\Delta_2}\right)_{L\in\mathcal{L}}=\left(f^L\right)_{L\in\mathcal{L}}$, normalized by setting $f^{1}=1$, equivalently $W_0 = V_\Delta$. Iterating Eq. \eqref{opew}, we can compute $L^*W_N$ for any creation operator $L\in\mathcal{L}_N$. On the other hand, $L^*W_N = \sum_{L'\in\mathcal{L}_N} S_{L,L'} f^{L'}V_\Delta$ where $S_{L,L'}$ is the Shapovalov form \eqref{shap}.
This leads to the linear system
\begin{align}
\forall L\in \mathcal{L}_N, \ \sum_{L'\in\mathcal{L}_N} S_{L,L'} f^{L'} = g^L \quad \text{with} \quad
g^{L_{-n_1}L_{-n_2}\cdots L_{-n_k}}_{\Delta,\Delta_2,\Delta_1} = \prod_{i=1}^k \theta_{\sum_{j=i}^k n_j, n_i}
\ .
\label{sgl}
\end{align}
Assuming the Shapovalov form is invertible at level $N$, the solution is
\begin{align}
\boxed{f^{\Delta,L}_{\Delta_1,\Delta_2} = \sum_{|L'|=|L|} S_{L,L'}^{-1}(\Delta) g^{L'}_{\Delta,\Delta_2,\Delta_1}}\ .
\label{fsg}
\end{align}
For example, we have $g^{L_{-1}} = \theta_{1,1}$, $g^{L_{-1}^2} = \theta_{1,1}\theta_{2,1}$ and $g^{L_{-2}}=\theta_{2,2}$ with
\begin{align}
\theta_{1,1} = \Delta+\Delta_1-\Delta_2 \quad , \quad \theta_{2,1} = \theta_{1,1}+1 \quad , \quad \theta_{2,2} = \Delta+2\Delta_2-\Delta_1 \ .
\end{align}
Inverting the Shapovalov matrix \eqref{sh012}, we deduce
\begin{subequations}
\label{fl12}
\begin{align}
f^{L_{-1}} & = \frac{\Delta+\Delta_1-\Delta_2}{2\Delta}\ ,
\label{fl1}
\\
\left[\begin{smallmatrix}
f^{L_{-1}^2} \\ f^{L_{-2}}
\end{smallmatrix}\right]
& =
\frac{1}{16(\Delta-\Delta_{(2,1)})(\Delta-\Delta_{(1,2)})}
\left[\begin{smallmatrix} 2+\frac{c}{4\Delta} & -3 \\ -3 & 4\Delta+2 \end{smallmatrix}\right]
\left[\begin{smallmatrix} (\Delta+\Delta_1-\Delta_2)(\Delta+\Delta_1-\Delta_2+1) \\ \Delta+2\Delta_1-\Delta_2 \end{smallmatrix}\right] \ .
\end{align}
\end{subequations}
The solution $\left(f^L\right)_{L\in\mathcal{L}}$ of the OPE Ward identities is unique if and only if the Shapovalov matrix is invertible, i.e. if and only if there is no null vector in the Verma module $\mathcal{V}_\Delta$. A null vector $L_{\langle r,s\rangle} V_{\Delta_{(r,s)}}$ at level $rs$ manifests itself as a simple pole at $\Delta=\Delta_{(r,s)}$ of the vectors $(f^L)_{L\in\mathcal{L}_N}$ for any $N\geq rs$.
To summarize, the contributions of $V_\Delta$ and its descendants in an OPE of 2 primary fields read
\begin{align}
\boxed{V_{\Delta_1}(z_1)V_{\Delta_2}(z_2) \supset C_{\Delta_1,\Delta_2}^{\Delta} z_{12}^{\Delta-\Delta_1-\Delta_2} \Bigg(V_\Delta(z_2) +\sum_{L\in \mathcal{L}\backslash\{1\}}z_{12}^{|L|} f^{\Delta,L}_{\Delta_1,\Delta_2} LV_\Delta(z_2)\Bigg)}
\ ,
\label{prope}
\end{align}
where the \myindex{OPE structure constant}\index{OPE!---structure constant}\index{structure constant!OPE---} $C^\Delta_{\Delta_1,\Delta_2}$ is left undetermined by conformal symmetry, while the dependence on $z_1,z_2$ and the coefficients $f^{\Delta,L}_{\Delta_1,\Delta_2}$ are universal quantities that are in principle known.
In the case of the OPE coefficients $C^{\Delta,L}_{\Delta_1,\Delta_2}(z_1,z_2)$, we have seen that the primary field determines the contributions of its descendant fields. This will also be true in correlation functions, as a result of Virasoro symmetry. However, this would not necessarily hold if we had a larger chiral symmetry algebra such as a W-algebra.
The form of the dependence on $z_1,z_2$ provides an a posteriori justification for our assumption in Section \ref{sec:hwr} that $L_0$-eigenvalues be bounded from below. We have indeed seen that an $L_0$-eigenvector $LV_\Delta$ of dimension $\Delta+|L|$ comes with a factor $z_{12}^{\Delta+|L|}$. If $L_0$-eigenvalues were not bounded from below, the sum over states could not converge in a neighbourhood of $z_1=z_2$.
\subsubsection{Fusion rules and fusion products}\label{sec:dope}
Because of the poles in OPE coefficients, there are constraints on the fields that can appear in OPEs. Let us focus on the pole at $\Delta_{(1,1)}=0$. There are 2 types of primary fields of dimension $0$: the non-degenerate field $V_0$, which generates the Verma module $\mathcal{V}_0$, and the degenerate field $V^d_{\langle 1,1\rangle}$, which obeys the null vector equation $L_{-1}V^d_{\langle 1,1\rangle}=\partial V^d_{\langle 1,1\rangle}=0$ and generates the degenerate representation $\mathcal{R}^d_{\langle 1,1\rangle}$ \eqref{r11}.
\begin{itemize}
\item Can $V_0$ appear in the OPE $V_{\Delta_1}V_{\Delta_2}$? The pole of $f^{L_{-1}}$ \eqref{fl1} means we cannot impose the normalization condition $f^1=1$: we in fact have $f^1=0$ while $f^{L_{-1}}\neq 0$. Therefore, the non-vanishing null vector $L_{-1}V_0$ may appear in the OPE, together with its descendants, while $V_0$ itself and its other descendants give vanishing contributions. This situation is better interpreted as the appearance of $V_1$, not $V_0$.
\item In the case of $V^d_{\langle 1,1\rangle}$, the null vector vanishes, and no non-vanishing descendant of $V^d_{\langle 1,1\rangle}$ may appear --- unless $g^{L_{-1}}=0$ in Eq. \eqref{sgl}, which happens if $\Delta_1=\Delta_2$.
\end{itemize}
On the other hand, the OPE $V_0V_{\Delta_2}$ is just a special case of the OPE \eqref{prope}.
But the OPE $V^d_{\langle 1,1\rangle}V_{\Delta_2}$ must obey
\begin{align}
0 = \frac{\partial}{\partial z_1} V^d_{\langle 1,1\rangle}(z_1)V_{\Delta_2}(z_2) \supset C_{\langle 1,1\rangle,\Delta_2}^{\Delta} \frac{\partial}{\partial z_1} z_{12}^{\Delta-\Delta_2} \big(V_\Delta(z_2) +\cdots \big) \ ,
\end{align}
therefore $V_\Delta$ may appear only provided $\Delta=\Delta_2$. This agrees with the remark (in Section \ref{sec:ope}) that $V^d_{\langle 1,1\rangle}$ is proportional to the identity field.
Let us introduce the \myindex{fusion rules}\index{fusion!---rule} as a convenient notation for writing constraints on OPEs, by omitting everything but the primary fields:
\begin{align}
V^d_{\langle 1,1\rangle} V_{\Delta} = V_\Delta \quad , \quad V_{\Delta_1} V_{\Delta_2} \ni V^d_{\langle 1,1\rangle} \implies \Delta_1=\Delta_2\ .
\end{align}
This may be reformulated in terms of
the \myindex{fusion product}\index{fusion!---product} of representations of the Virasoro algebra:
\begin{align}
\mathcal{R}^d_{\langle 1,1\rangle} \times \mathcal{V}_\Delta = \mathcal{V}_\Delta \quad , \quad \dim\text{Hom}\left(\mathcal{V}_{\Delta_1}\times \mathcal{V}_{\Delta_2}, \mathcal{R}^d_{\langle 1,1\rangle}\right) = \delta_{\Delta_1,\Delta_2}\ ,
\label{rtv}
\end{align}
where $\text{Hom}(\mathcal{R},\mathcal{R}')$ is the space of linear maps from $\mathcal{R}$ to $\mathcal{R}'$ that commute with the action of the Virasoro algebra.
The fusion product can be algebraically defined \cite{gab99}. It differs from the tensor product, in particular its definition involves the fields' positions, although the resulting product does not. Just like OPEs, the fusion product is commutative and associative. Moreover, as we just saw in the case of Verma modules, we have an isomorphism
\begin{align}
\text{Hom}\left(\mathcal{R},\mathcal{R}'\right) \simeq \text{Hom}\left(\mathcal{R}\times\mathcal{R}',\mathcal{R}^d_{\langle 1,1\rangle}\right)\ ,
\end{align}
which together with associativity implies
\begin{align}
\text{Hom}\left(\mathcal{R}\times\mathcal{R}',\mathcal{R}''\right) \simeq \text{Hom}\left(\mathcal{R}'\times\mathcal{R}'',\mathcal{R}\right)\ .
\label{hrr}
\end{align}
Given 3 representations $\mathcal{R},\mathcal{R}',\mathcal{R}''$, the number $\dim \text{Hom}\left(\mathcal{R}\times\mathcal{R}',\mathcal{R}''\right) \in \mathbb{N}$ is called the \myindex{fusion multiplicity}\index{fusion!---multiplicity} of $\mathcal{R}''$ in the fusion product $\mathcal{R}\times\mathcal{R}'$. A fusion multiplicity is called nontrivial if it is $\geq 2$.
\subsubsection{Fusion products of degenerate representations}
According to the isomorphism \eqref{hrr}, $V^d_{\langle 2,1\rangle} V_{\Delta_1} \ni V_{\Delta_2} \iff V_{\Delta_1}V_{\Delta_2} \ni V^d_{\langle 2,1\rangle} \iff $ the 2-dimensional linear system \eqref{sgl} at $N=2$ with $\Delta=\Delta_{(2,1)}$ has a solution, even though it is of rank 1. Explicitly, this condition reads
\begin{align}
\left(4\Delta_{(2,1)}+2\right) \left(\Delta_{(2,1)}+¨2\Delta_1-\Delta_2\right) = 3\left(\Delta_{(2,1)}+\Delta_1-\Delta_2\right)\left(\Delta_{(2,1)}+1+\Delta_1-\Delta_2\right)\ .
\end{align}
Using the expression \eqref{ars} for $\Delta_{(2,1)}$, and replacing conformal dimensions with momentums \eqref{dp}, this equation reduces to
\begin{align}
\prod_{\pm,\pm}\left(\tfrac{\beta}{2}\pm P_1\pm P_2\right) = 0\ .
\end{align}
This leads to the fusion rules of $V^d_{\langle 2,1\rangle}$, from which the fusion rules of $V^d_{\langle 1,2\rangle}$ are deduced by $\beta\to \beta^{-1}$. We write these fusion rules in terms of fusion products, where we index Verma modules by their momentums:
\begin{align}
\boxed{\mathcal{R}^d_{\langle 2,1\rangle}\times \mathcal{V}_P = \sum_\pm \mathcal{V}_{P\pm \frac{\beta}{2}}} \quad, \quad \boxed{\mathcal{R}^d_{\langle 1,2\rangle}\times \mathcal{V}_P = \sum_\pm \mathcal{V}_{P\pm \frac{1}{2\beta}}}\ .
\label{rvvp}
\end{align}
From these simple formulas, all the discrete spectrums of Section \ref{sec:sesc} will follow. For the moment, let us deduce the fusion products of higher degenerate representations $\mathcal{R}^d_{\langle r,s\rangle}$, using the associativity of the fusion product. We start with the remark that a highest-weight representation $\mathcal{R}$ is degenerate if and only if its fusion product with any Verma module $\mathcal{R}\times \mathcal{V}_P$ has finitely many terms:
\begin{itemize}
\item If $\mathcal{R}=\mathcal{R}^d_{\langle r,s\rangle}$ is degenerate, the condition that $V^d_{\langle r,s\rangle}$ may appear in the OPE $V_{\Delta_1}(z_1)V_{\Delta_2}(z_2)$ leads to polynomial equations of order $rs$ on $\Delta_1,\Delta_2$, just as we saw in the case $(r,s)=(2,1)$. These equations may be interpreted as conditions on residues of OPE coefficients,
\begin{align}
V_{\Delta_1}(z_1)V_{\Delta_2}(z_2) \supset V^d_{\langle r,s\rangle} \quad \implies \quad \forall L\in \mathcal{L},\ \underset{\Delta=\Delta_{(r,s)}}{\operatorname{Res}} f^{\Delta,L}_{\Delta_1,\Delta_2} = 0\ .
\end{align}
(If the residues vanish for $|L|=rs$, then they vanish for all $L\in\mathcal{L}$.)
\item If $\mathcal{R}$ is not degenerate, the corresponding primary field may appear in any OPE of the type $V_{\Delta_1}(z_1)V_{\Delta_2}(z_2)$, because the OPE Ward identities have a solution.
\end{itemize}
It follows that a fusion product of degenerate representations is itself a finite sum of degenerate representations.
For example, according to Eq. \eqref{rvvp}, the product $\mathcal{R}^d_{\langle 2,1\rangle}\times\mathcal{R}^d_{\langle 2,1\rangle}$ is the sum of 2 representations of momentums $\{P_{(2,1)} \pm \frac{\beta}{2}\} = \{P_{(1,1)},P_{(3,1)}\}$. These representations must be degenerate, therefore $\mathcal{R}^d_{\langle 2,1\rangle}\times\mathcal{R}^d_{\langle 2,1\rangle}=\mathcal{R}^d_{\langle 1,1\rangle}+\mathcal{R}^d_{\langle 3,1\rangle}$. By associativity of $\mathcal{R}^d_{\langle 2,1\rangle}\times\mathcal{R}^d_{\langle 2,1\rangle}\times \mathcal{V}_P$, we then deduce
$
\mathcal{R}^d_{\langle 3,1\rangle} \times \mathcal{V}_P = \mathcal{V}_{P-\beta}+\mathcal{V}_P+\mathcal{V}_{P+\beta}
$. Iterating this reasoning over $r$ or $s$, we obtain
\begin{align}
\mathcal{R}^d_{\langle 2,1\rangle}\times \mathcal{R}^d_{\langle r,s\rangle}
\underset{r\geq 2}{=} \sum_\pm \mathcal{R}^d_{\langle r\pm 1,s\rangle}
\quad , \quad
\mathcal{R}^d_{\langle 1,2\rangle}\times \mathcal{R}^d_{\langle r,s\rangle}
\underset{s\geq 2}{=}\sum_\pm \mathcal{R}^d_{\langle r,s\pm 1\rangle} \ .
\label{rrrs}
\end{align}
This allows us to find the fusion product of any degenerate representation with a Verma module, by iteration over $r$ and $s$:
\begin{align}
\boxed{\mathcal{R}^d_{\langle r,s\rangle}\times \mathcal{V}_P = \sum_{i=-\frac{r-1}{2}}^{\frac{r-1}{2}} \sum_{j=-\frac{s-1}{2}}^{\frac{s-1}{2}} \mathcal{V}_{P+i\beta +j\beta^{-1}}}\ ,
\label{rrsvp}
\end{align}
where the sums run by increments of $1$. Since this sum has $rs$ terms, the Verma module $\mathcal{V}_{\Delta_{(r,s)}}$ must have a null vector at level $rs$, a result that we stated in Section \ref{sec:nvvm}, and which we now derived using fusion products. Fusion can even be used to compute the null vectors themselves \cite{fms97}.
We can then find the fusion product $\mathcal{R}^d_{\langle r_1,s_1\rangle}\times \mathcal{R}^d_{\langle r_2,s_2\rangle}$, either by again iterating over $r_1$ and $s_1$, or by summing the degenerate representations whose momentums appear in both $\mathcal{R}^d_{\langle r_1,s_1\rangle}\times \mathcal{V}_{P_{(r_2,s_2)}}$ and $\mathcal{R}^d_{\langle r_2,s_2\rangle}\times \mathcal{V}_{P_{(r_1,s_1)}}$:
\begin{align}
\boxed{\mathcal{R}^d_{\langle r_1,s_1 \rangle} \times \mathcal{R}^d_{\langle r_2,s_2 \rangle} = \sum_{r\overset{2}{=}|r_1-r_2|+1}^{r_1+r_2-1}\ \sum_{s\overset{2}{=}|s_1-s_2|+1}^{s_1+s_2-1} \mathcal{R}^d_{\langle r,s \rangle}}\ ,
\label{rrsr}
\end{align}
where the sums now run by increments of $2$. The number of terms is $\min(r_1,r_2)\min(s_1,s_2)$. For example,
\begin{align}
\mathcal{R}^d_{\langle 5,2\rangle}\times \mathcal{R}^d_{\langle 3,3\rangle} = \mathcal{R}^d_{\langle 3,2\rangle} + \mathcal{R}^d_{\langle 5, 2\rangle} + \mathcal{R}^d_{\langle 7,2\rangle} +\mathcal{R}^d_{\langle 3,4\rangle}+\mathcal{R}^d_{\langle 5,4\rangle} + \mathcal{R}^d_{\langle 7,4\rangle} \ .
\end{align}
\subsection{Correlation functions}\label{sec:cor}
Correlation functions are observables of quantum field theory: quantities that we can in principle compute, and compare with experimental data. We will focus on $N$-point functions on the Riemann sphere: functions of $N$ distinct positions $z_1,z_2,\dots, z_N\in \overline{\mathbb{C}}$, which are written in terms of $N$ fields as
\begin{align}
\Big< V_1(z_1) V_2(z_2) \cdots V_N(z_N)\Big>\ .
\end{align}
Correlation functions are assumed to depend linearly on fields: in particular, this implies $\partial_{z_1}\left< V_1(z_1) \cdots \right> = \left< \partial_{z_1} V_1(z_1) \cdots \right>$.
All the assumptions and results of Section \ref{sec:fope} may be understood as constraints on correlation functions.
In 2d CFT, there are other observables that we will not consider, such as:
\begin{itemize}
\item Correlation functions on higher-genus Riemann surfaces, starting with the torus.
\item Correlation functions on Riemann surfaces with boundaries, starting with the disc.
\item Correlation functions in the presence of defect lines.
\item Entanglement entropy.
\end{itemize}
\subsubsection{Conformal symmetry constraints}\label{sec:csc}
In order to derive the conformal symmetry constraints for an $N$-point function of primary fields $Z$, it is convenient to consider the $N+1$-point function $Z(y)$ that also includes the energy-momentum tensor $T(y)$:
\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}
The function $Z(y)$ is meromorphic, with $N$ poles at $y=z_i$, where its behaviour is constrained by the OPE $T(y)V_{\Delta_i}(z_i)$ \eqref{tvd}. To fully determine $Z(y)$, it remains to determine its behaviour at $y=\infty$. Because our fields live on the Riemann sphere, they must be smooth at $y=\infty$: in the case of $T(y)$, we assume that this means
\begin{align}
T(y) \underset{y\to\infty} = O\left(\frac{1}{y^4}\right)\ .
\label{tinf}
\end{align}
This implies first of all that $\lim_{y\to \infty} Z(y)=0$, therefore $Z(y)$ is completely determined by its residues at $y=z_i$:
\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{zy}
\end{align}
Furthermore, Eq. \eqref{tinf} implies the vanishing of the coefficients of $y^{-1},y^{-2},y^{-3}$ in the Laurent expansion of $Z(y)$ near $y=\infty$, leading to the 3 \myindex{global Ward identities}\index{Ward identity!global---}
\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{gward}
\end{align}
Equivalently, the behaviour of $Z$ under global conformal transformations \eqref{abcd} must be
\begin{align}
\boxed{\left< \prod_{i=1}^N V_{\Delta_i}\left(\frac{az_i+b}{cz_i+d}\right) \right>
= \prod_{i=1}^N (cz_i +d)^{2\Delta_i} \left< \prod_{i=1}^N V_{\Delta_i}(z_i) \right>}\ .
\label{zgc}
\end{align}
The global Ward identities are valid not only for primary fields, but also for quasi-primary fields: $L_0$-eigenvectors that are annihilated by $L_1$ but not necessarily by $L_{n\geq 2}$.
In particular, applying Eq. \eqref{zgc} to the transformation $z\mapsto \frac{1}{z}$, we find $V_\Delta(z)\underset{z\to\infty}{=} O(z^{-2\Delta})$, which allows us to define
\begin{align}
V_\Delta(\infty) = \lim_{z\to\infty} z^{2\Delta}V_\Delta(z)\ .
\label{vdi}
\end{align}
Furthermore,
according to the OPE $T(y)T(z)$ \eqref{tt}, the energy-momentum tensor $T$ is a quasi-primary field of dimension $\Delta=2$, whose assumed behaviour at $\infty$ \eqref{tinf} is consistent with Eq. \eqref{vdi}.
Using Eq. \eqref{zgc} for a correlation function that involves $T$, we deduce a definition of descendant fields at $\infty$,
\begin{align}
L_n V_\Delta(\infty) = \frac{1}{2\pi i} \oint_\infty dy\ y^{1-n} T(y)V_\Delta(\infty)\ .
\label{lnia}
\end{align}
(Compare with $L_nV_\Delta(z)$ \eqref{lvtv}.)
We could alternatively use the transformation $z\mapsto \frac{1}{z-a}$ for any $a\in\mathbb{C}$, and obtain an $a$-dependent definition $ \frac{1}{2\pi i} \oint_\infty dy\ (y-a)^{1-n} T(y)V_\Delta(\infty)$.
We interpret $\frac{1}{z-a}$ as a \myindex{local coordinate}\index{local coordinate} near $z=\infty$. Correlation functions on general Riemann surfaces depend on choices of local coordinates, see \cite[Section 3]{br24}. In the case of the Riemann sphere, there is
a canonical local coordinate near any point $z\in\mathbb{C}$, but not near $z=\infty$. So we chose the local coordinate $\frac{1}{z}$ near $z=\infty$.
Let us solve the global Ward identities in the cases $N=1,2,3,4$. We have $3$ linear differential equations in the $N$ variables $z_i$, and we write their solutions up to $z_i$-independent prefactors:
\begin{itemize}
\item $\boxed{N=1}$\ : The identities amount to $\partial_{z_1} Z = \Delta_1 Z=0$, and their solution is
\begin{subequations}
\begin{align}
\Big<V_{\Delta_1}(z_1)\Big>\propto \delta_{\Delta_1,0}\ .
\end{align}
\item $\boxed{N=2}$\ : Again, there are more Ward identities than variables, leading to $Z\neq 0\implies \Delta_1=\Delta_2$, and to the solution
\begin{align}
\Big< V_{\Delta_1}(z_1)V_{\Delta_2}(z_2) \Big> \propto \delta_{\Delta_1,\Delta_2} z_{12}^{-2\Delta_1} \ .
\label{2pt}
\end{align}
\item $\boxed{N=3}$\ : With as many Ward identities as variables, we have a unique solution with no constraints on the conformal dimensions:
\begin{align}
\left< \prod_{i=1}^3 V_{\Delta_i}(z_i) \right> \propto z_{12}^{\Delta_3-\Delta_1-\Delta_2} z_{13}^{\Delta_2-\Delta_1-\Delta_3} z_{23}^{\Delta_1-\Delta_2-\Delta_3}\ .
\label{3pt}
\end{align}
\item $\boxed{N=4}$\ : The solution involves an arbitrary function $G(z)$ of the \myindex{cross-ratio}\index{cross-ratio} $z=\frac{z_{12}z_{34}}{z_{13}z_{24}}$:
\begin{align}
\left< \prod_{i=1}^4 V_{\Delta_i}(z_i) \right>
\propto z_{13}^{-2\Delta_1} z_{23}^{\Delta_1-\Delta_2-\Delta_3+\Delta_4} z_{24}^{-\Delta_1-\Delta_2+\Delta_3-\Delta_4} z_{34}^{\Delta_1+\Delta_2-\Delta_3-\Delta_4} G\left(\frac{z_{12}z_{34}}{z_{13}z_{24}}\right)\ .
\label{4pt}
\end{align}
\end{subequations}
A 4-point function is therefore completely determined by its dependence on $1$ of its $4$ variables. In particular, it can be convenient to set $3$ positions to the fixed values $(z_2,z_3,z_4)=(0,\infty,1)$, in which case
\begin{align}
\Big< V_{\Delta_1}(z) V_{\Delta_2}(0)V_{\Delta_3}(\infty)V_{\Delta_4}(1) \Big> \propto G(z)\ .
\end{align}
\end{itemize}
In addition to global Ward identities, there are \myindex{local Ward identities}\index{Ward identity!local---} , which follow from $\oint_\infty dy\ \epsilon(y)Z(y)=0$, where $\epsilon(y)$ is a meromorphic function such that $\epsilon(y)\underset{y\to\infty}{=}O(y^2)$, with poles at $y=z_i$. Local Ward identities allow us to determine correlation functions of descendant fields from the corresponding correlation functions of primary fields. In the case $\epsilon(y) = \frac{1}{(y-z_1)^{n-1}}$ with $n\geq -1$, using the expressions \eqref{zy} for $Z(y)$ and \eqref{lvtv} for $L_{-n}V_{\Delta_1}(z_1)$, we obtain the Ward identity
\begin{align}
\left<L_{-n}V_{\Delta_1}(z_1)\prod_{j=2}^N V_{\Delta_j}(z_j)\right> = \sum_{j=2}^N \left(-\frac{1}{z_{j1}^{n-1}}\frac{\partial}{\partial z_j} +\frac{n-1}{z_{j1}^n} \Delta_j\right) \left<\prod_{j=1}^N V_{\Delta_j}(z_j)\right>\ .
\label{lwi}
\end{align}
If $n\geq 2$ this is a local Ward identity, while for $n=-1,0,1$ we recover global Ward identities.
In particular, for any creation operator $L\in\mathcal{L}$ we can compute the ratio
\begin{align}
g^L_{\Delta_1,\Delta_2,\Delta_3} = \frac{\left< LV_{\Delta_1}(0) V_{\Delta_2}(\infty)V_{\Delta_3}(1)\right>}{\left< V_{\Delta_1}(0) V_{\Delta_2}(\infty)V_{\Delta_3}(1)\right>}\ .
\label{gl}
\end{align}
Let us show that this coincides with the quantity that appeared in OPE Ward identities \eqref{sgl}.
In the 3-point function $\left< L'V_\Delta(0)V_{\Delta_2}(\infty)V_{\Delta_1}(1)\right>\propto g^{L'}_{\Delta,\Delta_2,\Delta_1}$, we write the OPE $V_{\Delta_1}(1)V_{\Delta_2}(\infty)$ as a linear combination of descendant fields $LV_{\Delta}(\infty)$. Since our local coordinate near $\infty$ is $y=\frac{1}{z}$, we obtain a series in powers of $y_1-y_2=1$, which may seem large until we remember that the field $L'V_{\Delta}(0)$ is infinitely far away at $y=\infty$.
Assuming for simplicity $\left<V_\Delta(0) V_\Delta(\infty) \right>=1$,
we obtain
\begin{align}
g^{L'}_{\Delta,\Delta_2,\Delta_1}= \sum_{L\in\mathcal{L}} f^{\Delta,L}_{\Delta_1,\Delta_2} \Big<L'V_\Delta(0) LV_\Delta(\infty) \Big> \ .
\label{gfr}
\end{align}
By the definition \eqref{lnia} of descendant fields at $\infty$, the 2-point function coincides with the Shapovalov form,
\begin{align}
\boxed{S_{L,L'}(\Delta) = \Big<L'V_\Delta(0) LV_\Delta(\infty) \Big> = \Big<L^*L'V_\Delta(0) V_\Delta(\infty) \Big>} \ ,
\label{sll}
\end{align}
which vanishes if $|L|>|L'|$ because $V_\Delta$ is primary, and also if $|L|<|L'|$ because $LV_\Delta$ is at $\infty$. So we recover the expression \eqref{sgl} of $g^L$ in terms of OPE coefficients and the Shapovalov form.
\subsubsection{Single-valuedness and chiral factorization}
We assume that correlation functions are \myindex{single-valued}\index{single-valuedness} on the Riemann sphere. Equivalently, we assume that any 2 fields are \myindex{mutually local}\index{mutual locality}, i.e. that they have trivial monodromies around each other: $V_1(ze^{2\pi i})V_2(0)=V_1(z)V_2(0)$.
The correlation functions of Section \ref{sec:csc} are of the form $z^\Delta$, which is multivalued unless $\Delta\in\mathbb{Z}$. Single-valued functions like $|z|^{2\Delta}=z^\Delta\bar{z}^\Delta$ are not written in terms of powers of $z$, but also involve $\bar z$. In the literature, the notation $f(z,\bar z)$ is sometimes used for arbitrary functions of $z$, with $f(z)$ reserved to locally holomorphic functions, i.e. functions such that $\frac{\partial}{\partial \bar z} f(z) =0$. We will not use the notation $f(z,\bar z)$, which is redundant since $\bar z$ is itself a function of $z$.
The generators $\ell_n$ \eqref{lpz} of local conformal transformations are only valid when acting on locally holomorphic functions: for more general functions, the generators are
\begin{align}
\ell_n +\bar{\ell}_n \quad , \quad i(\ell_n -\bar \ell_n)\ .
\label{epbe}
\end{align}
In fact we want the algebra of conformal transformations to act on a complex vector space: the space of complex-valued functions on $\overline{\mathbb{C}}$ for the Witt algebra, the space of states of the CFT for the Virasoro algebra. We should therefore complexify our symmetry algebra, and we obtain 2 copies of the Virasoro algebra: a \myindex{left-moving} Virasoro algebra\index{left-moving} $\mathfrak{V}$ with generators $L_n$, and a \myindex{right-moving} Virasoro algebra\index{right-moving} $\bar{\mathfrak{V}}$ with generators $\bar{L}_n$. In the case of the algebra of global conformal transformations, we start with the 6-dimensional real Lie algebra $\mathfrak{sl}_2(\mathbb{C})$, with generators \eqref{epbe} with $n=-1,0,1$, and the complexification yields $\mathfrak{sl}_2(\mathbb{C})^\mathbb{C} = \mathfrak{sl}_2(\mathbb{C})\times \mathfrak{sl}_2(\mathbb{C})$.
In Sections \ref{sec:vir} and \ref{sec:fope}, we have been doing \myindex{chiral CFT}\index{CFT!chiral---}, which means considering $\mathfrak{V}$ only. In a full CFT, the symmetry algebra is the \myindex{conformal algebra}\index{conformal!---algebra} $\mathfrak{V}\times \bar{\mathfrak{V}}$, where for simplicity we assume that the left and right central charges are equal. When it comes to representations, Verma modules of $\mathfrak{V}\times \bar{\mathfrak{V}}$ are simply $\mathcal{V}_\Delta\otimes \bar{\mathcal{V}}_{\bar \Delta}$, where we write a bar over representations of $\bar{\mathfrak{V}}$. There also exist representations of $\mathfrak{V}\times \bar{\mathfrak{V}}$ that are not factorized, in particular logarithmic representations \cite{nr20}.
Fields now belong to representations of $\mathfrak{V}\times \bar{\mathfrak{V}}$, and they obey $\frac{\partial}{\partial \bar z} V(z) = \bar L_{-1} V(z)$, which is the antiholomorphic version of Eq. \eqref{pvlv}. This allows us to define a right-moving energy-momentum tensor $\bar T(y) = \sum_{n\in\mathbb{Z}} \frac{\bar L_n^{(z)}}{(\bar y-\bar z)^{n+2}}$. While $T$ is locally holomorphic, $\bar T$ is locally antiholomorphic:
\begin{align}
\partial_{\bar z} T(z) = 0 \quad , \quad \partial_z \bar T(z) = 0 \ .
\end{align}
Let $V_{\Delta,\bar\Delta}(z)$ be a field that is primary with respect to the conformal algebra, with the left dimension $\Delta$ and the right dimension $\bar\Delta$. The difference of these dimensions is called the \myindex{conformal spin}\index{conformal!---spin}
\begin{align}
S = \Delta -\bar\Delta\ .
\label{sdd}
\end{align}
According to Eq. \eqref{zgc},
under a rotation $z\mapsto e^{i\theta}z$, corresponding to a global conformal transformation $\left(\begin{smallmatrix} a & b \\ c & d\end{smallmatrix}\right)= \left(\begin{smallmatrix} e^{\frac{i}{2}\theta} & 0 \\ 0 & e^{-\frac{i}{2}\theta}\end{smallmatrix}\right)$, a primary field transforms as
\begin{align}
V_{\Delta,\bar\Delta}(z) \to e^{iS\theta} V_{\Delta,\bar\Delta}\left(e^{i\theta} z\right) \ .
\label{veis}
\end{align}
In order to write correlation functions, let us define the \myindex{modulus squared notation}\index{modulus squared notation} for functions of the central charge, conformal dimensions, and positions,
\begin{align}
\left| f(c,\Delta, z)\right|^2 = f(c,\Delta, z)f(c,\bar\Delta,\bar z)\ ,
\label{fdz2}
\end{align}
where $z,\bar z$ are complex conjugates, while $\Delta,\bar\Delta$ are generally not.
For simplicity we write $V_i$ for a primary field of dimensions $\Delta_i,\bar\Delta_i$.
The solution of the left- and right-moving global Ward identities for 2-point functions is the modulus squared of Eq. \eqref{2pt},
\begin{align}
\boxed{ \Big< V_1(z_1)V_2(z_2) \Big> = B_{12} \left|\delta_{\Delta_1,\Delta_2} z_{12}^{-2\Delta_1}\right|^2} \ ,
\label{2ptb}
\end{align}
where the constant, $z_i$-independent coefficient $B_{12}$ is now called the \myindex{2-point structure constant}\index{structure constant!2-point---}. Similarly, 3-point functions are given by the modulus squared of Eq. \eqref{3pt},
\begin{align}
\boxed{ \Big< V_1(z_1)V_2(z_2)V_3(z_3) \Big> = C_{123} \left|z_{12}^{\Delta_3-\Delta_1-\Delta_2} z_{13}^{\Delta_2-\Delta_1-\Delta_3} z_{23}^{\Delta_1-\Delta_2-\Delta_3}\right|^2 }\ ,
\label{3ptc}
\end{align}
where $C_{123}$ is the \myindex{3-point structure constant}\index{structure constant!3-point---}. Single-valuedness of the 2-point and 3-point functions implies that spins obey