diff --git "a/19AzT4oBgHgl3EQfuP0B/content/tmp_files/2301.01686v1.pdf.txt" "b/19AzT4oBgHgl3EQfuP0B/content/tmp_files/2301.01686v1.pdf.txt" new file mode 100644--- /dev/null +++ "b/19AzT4oBgHgl3EQfuP0B/content/tmp_files/2301.01686v1.pdf.txt" @@ -0,0 +1,22345 @@ +String Field Theory – A Modern Introduction +Harold Erbin1* +*Center for Theoretical Physics +Massachusetts Institute of Technology, Cambridge, MA 02139, Usa +*Cea, List, Gif-sur-Yvette, F-91191, France +7th January 2023 +1erbin@mit.edu +arXiv:2301.01686v1 [hep-th] 4 Jan 2023 + +Abstract +This book provides an introduction to string field theory (SFT). String theory is usually +formulated in the worldsheet formalism, which describes a single string (first-quantization). +While this approach is intuitive and could be pushed far due to the exceptional properties of +two-dimensional theories, it becomes cumbersome for some questions or even fails at a more +fundamental level. These motivations have led to the development of SFT, a description of +string theory using the field theory formalism (second-quantization). As a field theory, SFT +provides a rigorous and constructive formulation of string theory. +The main objective is to construct the closed bosonic SFT and to explain how to assess +the consistency of string theory with it. The accent is put on providing the reader with +the foundations, conceptual understanding and intuition of what SFT is. After reading this +book, they should be able to study the applications from the literature. +The book is organized in two parts. The first part reviews the topics of the worldsheet +theory that are necessary to build SFT (worldsheet path integral, CFT and BRST quant- +ization). The second part starts by introducing general concepts of SFT from the BRST +quantization. Then, it introduces off-shell string amplitudes before providing a Feynman +diagrams interpretation from which the building blocks of SFT are extracted. After con- +structing the closed SFT, it is used to outline the proofs of several important consistency +properties, such as background independence, unitarity and crossing symmetry. Finally, the +generalization to the superstring is also discussed. +This book grew up from lecture notes for a course given at the Ludwig-Maximilians- +Universität LMU (winter semesters 2017–2018 and 2018–2019). The current document is +the draft of the manuscript published by Springer. + +Preface +This book grew up from lectures delivered within the Elite Master Program “Theoretical +and Mathematical Physics” from the Ludwig-Maximilians-Universität during the winter +semesters 2017–2018 and 2018–2019. +The main focus of this book is the closed bosonic string field theory (SFT). While there +are many resources available for the open bosonic SFT, a single review [71] has been written +since the final construction of the bosonic closed SFT by Zwiebach [262]. For this reason, +it makes sense to provide a modern and extensive study. Moreover, the usual approach to +open SFT focuses on the cubic theory, which is so special that it is difficult to generalize the +techniques to other SFTs. Finally, closed strings are arguably more fundamental than open +strings because they are always present since they describe gravity, which further motivates +my choice. However, the reader should not take this focus as denying the major achievements +and the beauty of the open SFT; reading this book should provide most of the tools needed +to feel comfortable also with this theory. +While part of the original goal of SFT is to provide a non-perturbative definition of string +theory and to address important questions such as classifying consistent string backgrounds +or understanding dualities, no progress on this front has been achieved so far. Hence, there +is still much to understand and the recent surge of developments provide a new chance +to deepen our understanding of closed SFT. For example, several consistency properties of +string theory have been proven rigorously using SFT. Moreover, the recent construction of +the open-closed superstring field theory [165] together with earlier works [42, 218, 262, 264] +show that all types of string theories can be recast as a SFT. This is why, I believe, it is a +good time to provide a complete book on SFT. +The goal of this book is to offer a self-contained description of SFT and all the tools +necessary to build it. The emphasis is on describing the concepts behind SFT and to make +the reader build intuitions on what it means. +For this reason, there are relatively few +applications. +The reader is assumed to have some knowledge of QFT, and a basic knowledge of CFT +and string theory (classical string, Nambu–Goto action, light-cone and old-covariant quant- +izations). +Organization +The text is organized on three levels: the main content (augmented with examples), compu- +tations, and remarks. The latter two levels can be omitted in a first lecture. The examples, +computations and remarks are clearly separated from the text (respectively, by a half-box +on the left and bottom, by a vertical line on the left, and by italics) to help the navigation. +Many computations have been set aside from the main text to avoid breaking the flow +and to provide the reader with the opportunity to check by themself first. In some occasions, +computations are postponed well below the corresponding formula to gather similar compu- +tations or to avoid breaking an argument. While the derivations contain more details than +2 + +usual textbooks and may look pedantic to the expert, I think it is useful for students and +newcomers to have complete references where to check each step. This is even more the case +when there are many different conventions in the literature. The remarks are not directly +relevant to the core of the text but they make connections with other parts or topics. The +goal is to broaden the perspectives of the main text. +General references can be found at the end of each chapter to avoid overloading the +text. In-text references are reserved for specific points or explicit quotations (of a formula, a +discussion, a proof, etc.). I did not try to be exhaustive in the citations and I have certainly +missed important references: this should be imputed to my lack of familiarity with them +and not to their value. +This text is a preprint of the textbook [64] and is reproduced with permission of Springer. +My plan is to frequently update the draft of this book with new content. The last version +can be accessed on my professional webpage, currently located at: +http://www.lpthe.jussieu.fr/~erbin/ +Acknowledgements +I have started to learn string field theory at Hri by attending lectures from Ashoke Sen. +Since then, I have benefited from collaboration and many insightful discussions with him. +Following his lectures have been much helpful in building an intuition that cannot be found +in papers or reviews on the topic. Through this book, I hope being able to make some of +these insights more accessible. +I am particularly grateful to Ivo Sachs who proposed me to teach this course and to +Michael Haack for continuous support and help with the organization, and to both of them +for many interesting discussions during the two years I have spent at Lmu. Moreover, I have +been very lucky to be assigned an excellent tutor for this course, Christoph Chiaffrino. After +providing him with the topic and few references, Christoph has prepared all the tutorials +and the corrections autonomously. His help brought a lot to the course. +I am particularly obliged to all the students who have taken this course at Lmu for +many interesting discussions and comments: Enrico Andriolo, Hrólfur Ásmundsson, Daniel +Bockisch, Fabrizio Cordonnier, Julian Freigang, Wilfried Kaase, Andriana Makridou, Pouria +Mazloumi, Daniel Panea, Martin Rojo. +I am also grateful to all the string theory community for many exchanges. For discussions +related to the topics of this book, I would like to thank more particularly: Costas Bachas, +Adel Bilal, Subhroneel Chakrabarti, Atish Dabholkar, Benoit Douçot, Ted Erler, Dileep +Jatkar, Carlo Maccaferri, Juan Maldacena, Yuji Okawa, Sylvain Ribault, Raoul Santachiara, +Martin Schnabl, Dimitri Skliros, Jakub Vošmera. I have received a lot of feedback during +the different stages of writing this book, and I am obliged to all the colleagues who sent me +feedback. +I am thankful to my colleagues at Lmu for providing a warm and stimulating envir- +onment, with special thanks to Livia Ferro for many discussions around coffee. Moreover, +the encouragements and advice from Oleg Andreev and Erik Plauschinn have been strong +incentives for publishing this book. +The editorial process at Springer has been very smooth. +I would also like to thank +Christian Caron and Lisa Scalone for their help and efficiency during the publishing process. +I am also indebted to Stefan Theisen for having supported the publication at Springer and +for numerous comments and corrections on the draft. +Most of this book was written at the Ludwig–Maximilians–Universität (Lmu, Munich, +Germany) where I was supported by a Carl Friedrich von Siemens Research Fellowship of the +Alexander von Humboldt Foundation. The final stage has been completed at the University +3 + +of Turin (Italy). My research is currently funded by the European Union’s Horizon 2020 +research and innovation program under the Marie Skłodowska-Curie grant agreement No +891169. +Finally, writing this book would have been more difficult without the continuous and +loving support from Corinne. +November 2020 +Harold Erbin +4 + +Contents +Preface +2 +1 +Introduction +11 +1.1 +Strings, a distinguished theory +. . . . . . . . . . . . . . . . . . . . . . . . . . +11 +1.2 +String theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +14 +1.2.1 +Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +14 +1.2.2 +Classification of superstring theories . . . . . . . . . . . . . . . . . . . +18 +1.2.3 +Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +20 +1.3 +String field theory +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +23 +1.3.1 +From the worldsheet to field theory . . . . . . . . . . . . . . . . . . . . +23 +1.3.2 +String field action +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . +25 +1.3.3 +Expression with spacetime fields +. . . . . . . . . . . . . . . . . . . . . +25 +1.3.4 +Applications +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +26 +1.4 +Suggested readings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +27 +I +Worldsheet theory +28 +2 +Worldsheet path integral: vacuum amplitudes +29 +2.1 +Worldsheet action and symmetries . . . . . . . . . . . . . . . . . . . . . . . . +29 +2.2 +Path integral +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +34 +2.3 +Faddeev–Popov gauge fixing . . . . . . . . . . . . . . . . . . . . . . . . . . . . +36 +2.3.1 +Metrics on Riemann surfaces +. . . . . . . . . . . . . . . . . . . . . . . +37 +2.3.2 +Reparametrizations and analysis of P1 . . . . . . . . . . . . . . . . . . +42 +2.3.3 +Weyl transformations and quantum anomalies . . . . . . . . . . . . . . +47 +2.3.4 +Ambiguities, ultralocality and cosmological constant . . . . . . . . . . +48 +2.3.5 +Gauge-fixed path integral . . . . . . . . . . . . . . . . . . . . . . . . . +49 +2.4 +Ghost action +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +51 +2.4.1 +Actions and equations of motion . . . . . . . . . . . . . . . . . . . . . +51 +2.4.2 +Weyl ghost +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +52 +2.4.3 +Zero-modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +54 +2.5 +Normalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +55 +2.6 +Summary +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +56 +2.7 +Suggested readings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +56 +3 +Worldsheet path integral: scattering amplitudes +58 +3.1 +Scattering amplitudes on moduli space . . . . . . . . . . . . . . . . . . . . . . +58 +3.1.1 +Vertex operators and path integral . . . . . . . . . . . . . . . . . . . . +58 +3.1.2 +Gauge fixing: general case . . . . . . . . . . . . . . . . . . . . . . . . . +61 +3.1.3 +Gauge fixing: 2-point amplitude +. . . . . . . . . . . . . . . . . . . . . +64 +5 + +3.2 +BRST quantization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +67 +3.2.1 +BRST symmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +68 +3.2.2 +BRST cohomology and physical states . . . . . . . . . . . . . . . . . . +69 +3.3 +Summary +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +71 +3.4 +Suggested readings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +71 +4 +Worldsheet path integral: complex coordinates +72 +4.1 +Geometry of complex manifolds . . . . . . . . . . . . . . . . . . . . . . . . . . +72 +4.2 +Complex representation of path integral . . . . . . . . . . . . . . . . . . . . . +75 +4.3 +Summary +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +77 +4.4 +Suggested readings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +77 +5 +Conformal symmetry in D dimensions +78 +5.1 +CFT on a general manifold +. . . . . . . . . . . . . . . . . . . . . . . . . . . . +78 +5.2 +CFT on Minkowski space +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . +79 +5.3 +Suggested readings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +80 +6 +Conformal field theory on the plane +81 +6.1 +The Riemann sphere . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +81 +6.1.1 +Map to the complex plane . . . . . . . . . . . . . . . . . . . . . . . . . +81 +6.1.2 +Relation to the cylinder – string theory +. . . . . . . . . . . . . . . . . +83 +6.2 +Classical CFTs +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +84 +6.2.1 +Witt conformal algebra +. . . . . . . . . . . . . . . . . . . . . . . . . . +85 +6.2.2 +PSL(2, C) conformal group +. . . . . . . . . . . . . . . . . . . . . . . . +86 +6.2.3 +Definition of a CFT +. . . . . . . . . . . . . . . . . . . . . . . . . . . . +88 +6.3 +Quantum CFTs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +89 +6.3.1 +Virasoro algebra +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +90 +6.3.2 +Correlation functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . +90 +6.4 +Operator formalism and radial quantization . . . . . . . . . . . . . . . . . . . +92 +6.4.1 +Radial ordering and commutators +. . . . . . . . . . . . . . . . . . . . +92 +6.4.2 +Operator product expansions . . . . . . . . . . . . . . . . . . . . . . . +94 +6.4.3 +Hermitian and BPZ conjugation +. . . . . . . . . . . . . . . . . . . . . +96 +6.4.4 +Mode expansion +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +97 +6.4.5 +Hilbert space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100 +6.4.6 +CFT on the cylinder . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105 +6.5 +Suggested readings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106 +7 +CFT systems +107 +7.1 +Free scalar . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 +7.1.1 +Covariant action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 +7.1.2 +Action on the complex plane +. . . . . . . . . . . . . . . . . . . . . . . 109 +7.1.3 +OPE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111 +7.1.4 +Mode expansions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113 +7.1.5 +Commutators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 +7.1.6 +Hilbert space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117 +7.1.7 +Euclidean and BPZ conjugates . . . . . . . . . . . . . . . . . . . . . . 118 +7.2 +First-order bc ghost system +. . . . . . . . . . . . . . . . . . . . . . . . . . . . 119 +7.2.1 +Covariant action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119 +7.2.2 +Action on the complex plane +. . . . . . . . . . . . . . . . . . . . . . . 119 +7.2.3 +OPE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121 +7.2.4 +Mode expansions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123 +7.2.5 +Commutators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125 +6 + +7.2.6 +Hilbert space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126 +7.2.7 +Euclidean and BPZ conjugates . . . . . . . . . . . . . . . . . . . . . . 132 +7.2.8 +Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133 +7.3 +Suggested readings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133 +8 +BRST quantization +134 +8.1 +BRST for reparametrization invariance . . . . . . . . . . . . . . . . . . . . . . 134 +8.2 +BRST in the CFT formalism +. . . . . . . . . . . . . . . . . . . . . . . . . . . 135 +8.2.1 +OPE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135 +8.2.2 +Mode expansions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136 +8.2.3 +Commutators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137 +8.3 +BRST cohomology: two flat directions . . . . . . . . . . . . . . . . . . . . . . 138 +8.3.1 +Conditions on the states . . . . . . . . . . . . . . . . . . . . . . . . . . 139 +8.3.2 +Relative cohomology . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141 +8.3.3 +Absolute cohomology, states and no-ghost theorem . . . . . . . . . . . 146 +8.3.4 +Cohomology for holomorphic and anti-holomorphic sectors . . . . . . . 147 +8.4 +Summary +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148 +8.5 +Suggested readings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148 +II +String field theory +149 +9 +String field +150 +9.1 +Field functional . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150 +9.2 +Field expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151 +9.3 +Summary +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152 +9.4 +Suggested readings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153 +10 Free BRST string field theory +154 +10.1 Classical action for the open string . . . . . . . . . . . . . . . . . . . . . . . . 154 +10.1.1 Warm-up: point-particle . . . . . . . . . . . . . . . . . . . . . . . . . . 154 +10.1.2 Open string action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155 +10.1.3 Gauge invariance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157 +10.1.4 Siegel gauge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158 +10.2 Open string field expansion, parity and ghost number +. . . . . . . . . . . . . 160 +10.3 Path integral quantization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162 +10.3.1 Tentative Faddeev–Popov gauge fixing . . . . . . . . . . . . . . . . . . 162 +10.3.2 Tower of ghosts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164 +10.4 Spacetime action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165 +10.5 Closed string +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167 +10.6 Summary +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169 +10.7 Suggested readings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170 +11 Introduction to off-shell string theory +171 +11.1 Motivations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171 +11.1.1 3-point function +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171 +11.1.2 4-point function +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173 +11.2 Off-shell states +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175 +11.3 Off-shell amplitudes +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177 +11.3.1 Amplitudes from the marked moduli space +. . . . . . . . . . . . . . . 178 +11.3.2 Local coordinates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179 +11.4 Suggested readings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181 +7 + +12 Geometry of moduli spaces and Riemann surfaces +182 +12.1 Parametrization of Pg,n +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182 +12.2 Tangent space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185 +12.3 Plumbing fixture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187 +12.3.1 Separating case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187 +12.3.2 Non-separating case +. . . . . . . . . . . . . . . . . . . . . . . . . . . . 190 +12.3.3 Decomposition of moduli spaces and degeneration limit +. . . . . . . . 191 +12.3.4 Stubs +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194 +12.4 Summary +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195 +12.5 Suggested readings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196 +13 Off-shell amplitudes +197 +13.1 Cotangent spaces and amplitudes . . . . . . . . . . . . . . . . . . . . . . . . . 197 +13.1.1 Construction of forms . . . . . . . . . . . . . . . . . . . . . . . . . . . 197 +13.1.2 Amplitudes and surface states . . . . . . . . . . . . . . . . . . . . . . . 199 +13.2 Properties of forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201 +13.2.1 Vanishing of forms with trivial vectors . . . . . . . . . . . . . . . . . . 202 +13.2.2 BRST identity +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203 +13.3 Properties of amplitudes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204 +13.3.1 Restriction to ˆPg,n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204 +13.3.2 Consequences of the BRST identity +. . . . . . . . . . . . . . . . . . . 206 +13.4 Suggested readings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207 +14 Amplitude factorization and Feynman diagrams +208 +14.1 Amplitude factorization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208 +14.1.1 Separating case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208 +14.1.2 Non-separating case +. . . . . . . . . . . . . . . . . . . . . . . . . . . . 212 +14.2 Feynman diagrams and Feynman rules . . . . . . . . . . . . . . . . . . . . . . 213 +14.2.1 Feynman graphs +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214 +14.2.2 Propagator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215 +14.2.3 Fundamental vertices . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217 +14.2.4 Stubs +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222 +14.2.5 1PI vertices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222 +14.3 Properties of fundamental vertices +. . . . . . . . . . . . . . . . . . . . . . . . 223 +14.3.1 String product +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223 +14.3.2 Feynman graph interpretation . . . . . . . . . . . . . . . . . . . . . . . 224 +14.4 Suggested readings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225 +15 Closed string field theory +226 +15.1 Closed string field expansion +. . . . . . . . . . . . . . . . . . . . . . . . . . . 226 +15.2 Gauge fixed theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227 +15.2.1 Kinetic term and propagator +. . . . . . . . . . . . . . . . . . . . . . . 227 +15.2.2 Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229 +15.2.3 Action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230 +15.3 Classical gauge invariant theory . . . . . . . . . . . . . . . . . . . . . . . . . . 232 +15.4 BV theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234 +15.5 1PI theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236 +15.6 Suggested readings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237 +8 + +16 Background independence +238 +16.1 The concept of background independence +. . . . . . . . . . . . . . . . . . . . 238 +16.2 Problem setup +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239 +16.3 Deformation of the CFT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240 +16.4 Expansion of the action +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242 +16.5 Relating the equations of motion . . . . . . . . . . . . . . . . . . . . . . . . . 242 +16.6 Idea of the proof . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243 +16.7 Suggested readings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244 +17 Superstring +245 +17.1 Worldsheet superstring theory . . . . . . . . . . . . . . . . . . . . . . . . . . . 245 +17.1.1 Heterotic worldsheet . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245 +17.1.2 Hilbert spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248 +17.2 Off-shell superstring amplitudes . . . . . . . . . . . . . . . . . . . . . . . . . . 249 +17.2.1 Amplitudes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 250 +17.2.2 Factorization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251 +17.2.3 Spurious poles +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252 +17.3 Superstring field theory +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255 +17.3.1 String field and propagator . . . . . . . . . . . . . . . . . . . . . . . . 256 +17.3.2 Constraint approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256 +17.3.3 Auxiliary field approach . . . . . . . . . . . . . . . . . . . . . . . . . . 257 +17.3.4 Large Hilbert space +. . . . . . . . . . . . . . . . . . . . . . . . . . . . 258 +17.4 Suggested readings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259 +18 Momentum-space SFT +260 +18.1 General form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260 +18.2 Generalized Wick rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262 +18.3 Suggested readings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264 +A Conventions +266 +A.1 Coordinates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266 +A.2 Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268 +A.3 QFT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268 +A.4 Curved space and gravity +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269 +A.5 List of symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270 +B Summary of important formulas +273 +B.1 +Complex analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273 +B.2 +QFT, curved spaces and gravity . . . . . . . . . . . . . . . . . . . . . . . . . . 273 +B.2.1 +Two dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274 +B.3 +Conformal field theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274 +B.3.1 +Complex plane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274 +B.3.2 +General properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275 +B.3.3 +Hermitian and BPZ conjugations . . . . . . . . . . . . . . . . . . . . . 276 +B.3.4 +Scalar field +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 276 +B.3.5 +Reparametrization ghosts . . . . . . . . . . . . . . . . . . . . . . . . . 277 +B.4 +Bosonic string . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 280 +9 + +C Quantum field theory +281 +C.1 Path integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281 +C.1.1 +Integration measure +. . . . . . . . . . . . . . . . . . . . . . . . . . . . 281 +C.1.2 +Field redefinitions +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283 +C.1.3 +Zero-modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283 +C.2 BRST quantization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 285 +C.3 BV formalism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 288 +C.3.1 +Properties of gauge algebra . . . . . . . . . . . . . . . . . . . . . . . . 289 +C.3.2 +Classical BV +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289 +C.3.3 +Quantum BV . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293 +C.4 Suggested readings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 294 +Bibliography +295 +Index +310 +10 + +Chapter 1 +Introduction +Abstract +In this chapter, we introduce the main motivations for studying string theory, +and why it is important to design a string field theory. After describing the central features +of string theory, we describe the most important concepts of the worldsheet formulation. +Then, we explain the reasons leading to string field theory (SFT) and outline the ideas +which will be discussed in the rest of the book. +1.1 +Strings, a distinguished theory +The first and simplest reason for considering theories of fundamental p-branes (fundamental +objects extended in p spatial dimensions) can be summarized by the following question: +“Why would Nature just make use of point-particles?” There is no a priori reason forbidding +the existence of fundamental extended objects and, according to Gell-Mann’s totalitarian +principle, “Everything not forbidden is compulsory.” If a consistent theory cannot be built +(after a reasonable amount of effort) or if it contradicts current theories (in their domains +of validity) and experiments, then one can support the claim that only point-particles exist. +On the other side, if such a theory can be built, it is of primary interest to understand it +deeper and to see if it can solve the current problems in high-energy theoretical physics. +The simplest case after the point particle is the string, so it makes sense to start with +it. It happens that a consistent theory of strings can be constructed, and that the latter +(in its supersymmetric version) contains all the necessary ingredients for a fully consistent +high-energy model:1 +• quantum gravity (quantization of general relativity plus higher-derivative corrections); +• grand unification (of matter, interactions and gravity); +• no divergences, UV finiteness (finite and renormalizable theory); +• fixed number of dimensions (26 = 25 + 1 for the bosonic string, 10 = 9 + 1 for the +supersymmetric version); +• existence of all possible branes; +• no dimensionless parameters and one dimensionful parameter (the string length ℓs). +It can be expected that a theory of fundamental strings (1-branes) occupies a distinguished +place among fundamental p-branes for the following reasons. +1There are also indications that a theory of membranes (2-branes) in 10+1 dimensions, called M-theory, +should exist. No direct and satisfactory description of the latter has been found and we will thus focus on +string theory in this book. +11 + +Figure 1.1: Locality of a particle interaction: two different observers always agree on the +interaction point and which parts of the worldline are 1- and 2-particle states. +Interaction non-locality +In a QFT of point particles, UV divergences arise because +interactions (defined as the place where the number and/or nature of the objects change) +are arbitrarily localized at a spacetime point. In Feynman graphs, such divergences can be +seen when the momentum of a loop becomes infinite (two vertices collide): this happens +when trying to concentrate an infinite amount of energy at a single point. However, these +divergences are expected to be reduced or absent in a field theory of extended objects: +whereas the interaction between particles is perfectly local in spacetime and agreed upon by +all observers (Figure 1.1), the spatial extension of branes makes the interactions non-local. +This means that two different observers will neither agree on the place of the interactions +(Figure 1.2), nor on the part of the diagram which describes one or two branes. +The string lies at the boundary between too much local and too much non-local: in any +given frame, the interaction is local in space, but not in spacetime. The reason is that a +string is one-dimensional and splits or joins along a point. For p > 1, the brane needs to +break/join along an extended spatial section, which looks non-local. +Another consequence of the non-locality is a drastic reduction of the possible interactions. +If an interaction is Lorentz invariant, Lorentz covariant objects can be attached at the vertex +(such as momentum or gamma matrices): this gives Lorentz invariants after contracting with +indices carried by the field. But, this is impossible if the interaction itself is non-local (and +thus not invariant): inserting a covariant object would break Lorentz invariance. +Brane degrees of freedom +The higher the number of spatial dimensions of a p-brane, the +more possibilities it has to fluctuate. As a consequence, it is expected that new divergences +appear as p increases due to the proliferations of the brane degrees of freedom. From the +worldvolume perspective, this is understood from the fact that the worldvolume theory +describes a field theory in (p + 1) dimensions, and UV divergences become worse as the +number of dimensions increase. +The limiting case happens for the string (p = 1) since +two-dimensional field theories are well-behaved in this respect (for example, any monomial +interaction for a scalar field is power-counting renormalizable). This can be explained by the +low-dimensionality of the momentum integration and by the enhancement of symmetries in +two dimensions. Hence, strings should display nice properties and are thus of special interest. +Worldvolume theory +The point-particle (0-brane) and the string (1-brane) are also re- +markable in another aspect: it is possible to construct a simple worldvolume field theory +(and the associated functional integral) in terms of a worldvolume metric. All components +of the latter are fixed by gauge symmetries (diffeomorphisms for the particle, diffeomorph- +isms and Weyl invariance for the string). This ensures the reparametrization invariance of +12 + +(a) Observers at rest and boosted. +(b) Observers close to the speed +of light moving in opposite dir- +ections. +The interactions are +widely separated in each case. +Figure 1.2: Non-locality of string interaction: two different observers see the interaction +happening at different places (denoted by the filled and empty circles) and they don’t agree +on which parts of the worldsheet are 1- and 2-string states (the litigation is denoted by the +grey zone). +the worldvolume without having to use a complicated action. Oppositely, the worldvolume +metric cannot be completely gauge fixed for p > 1. +Summary +As a conclusion, strings achieve an optimal balance between spacetime and +worldsheet divergences, as well as having a simple description with reparametrization in- +variance. +Since the construction of a field theory is difficult, it is natural to start with a worldsheet +theory and to study it in the first-quantization formalism, which will provide a guideline for +writing the field theory. In particular, this allows to access the physical states in a simple +way and to find other general properties of the theory. When it comes to the interactions +and scattering amplitudes, this approach may be hopeless in general since the topology of +the worldvolume needs to be specified by hand (describing the interaction process). In this +respect, the case of the string is again exceptional: because Riemann surfaces have been +classified and are well-understood, the arbitrariness is minimal. Combined with the tools of +conformal field theory, many computations can be performed. Moreover, since the modes +of vibrations of the strings provide all the necessary ingredients to describe the Standard +model, it is sufficient to consider only one string field (for one type of strings), instead +of the plethora found in point-particle field theory (one field for each particle). Similarly, +non-perturbative information (such as branes and dualities) could be found only due to the +specific properties of strings. +Coming back to the question which opened this section, higher-dimensional branes of all +the allowed dimensions naturally appear in string theory as bound states. Hence, even if +the worldvolume formulation of branes with p > 1 looks pathological2, string theory hints +towards another definition of these objects. +2Entering in the details would take us too far away from the main topic of this book. +Some of the +problems found when dealing with (p > 2)-branes are: how to define a Wick rotation for 3-manifolds, the +presence of Lorentz anomalies in target spacetime, problems with the spectrum, lack of renormalizability, +impossibility to gauge-fix the worldvolume metric [5–11, 41, 44, 45, 62, 127, 152, 156–158, 181, 193]. +13 + +1.2 +String theory +1.2.1 +Properties +The goal of this section is to give a general idea of string theory by introducing some concepts +and terminology. The reader not familiar with the points described in this section is advised +to follow in parallel some standard worldsheet string theory textbooks. +Worldsheet CFT +A string is characterized by its worldsheet field theory (Chapter 2).3 +The worldsheet is +parametrized by coordinates σa = (τ, σ). The simplest description is obtained by endowing +the worldsheet with a metric gab(σa) (a = 0, 1) and by adding a set of D scalar fields +Xµ(σa) living on the worldsheet (µ = 0, . . . , D − 1). The latter represents the position +of the string in the D-dimensional spacetime. From the classical equations of motion, the +metric gab is proportional to the metric induced on the worldsheet from its embedding in +spacetime. +More generally, one ensures that the worldsheet metric is non-dynamical by +imposing that the action is invariant under (worldsheet) diffeomorphisms and under Weyl +transformations (local rescalings of the metric). The consistency of these conditions at the +quantum level imposes that D = 26, and this number is called the critical dimension. Gauge +fixing the symmetries, and thus the metric, leads to the conformal invariance of the resulting +worldsheet field theory: a conformal field theory (CFT) is a field theory (possibly on a curved +background) in which only angles and not distances can be measured (Chapters 5 to 7). +This simplifies greatly the analysis since the two-dimensional conformal algebra (called the +Virasoro algebra) is infinite-dimensional. +CFTs more general than D free scalar fields can be considered: fields taking non-compact +values are interpreted as non-compact dimensions while compact or Grassmann-odd fields +are interpreted as compact dimensions or internal structure, like the spin. +While the light-cone quantization allows to find quickly the states of the theory, the +simplest covariant method is the BRST quantization (Chapter 8). It introduces ghosts (and +superghosts) associated to the gauge fixing of diffeomorphisms (and local supersymmetry). +These (super)ghosts form a CFT which is universal (independent of the matter CFT). +The trajectory of the string is denoted by xc(τ, σ). It begins and ends respectively at the +geometric shapes parametrized by xc(τi, σ) = xi(σ) and by xc(τf, σ) = xf(σ). Note that the +coordinate system on the worldsheet itself is arbitrary. The spatial section of a string can be +topologically closed (circle) or open (line) (Figure 1.3), leading to cylindrical or rectangular +worldsheets as illustrated in Figures 1.4 and 1.5. To each topology is associated different +boundary conditions and types of strings: +• closed: periodic and anti-periodic boundary conditions; +• open: Dirichlet and Neumann boundary conditions. +While a closed string theory is consistent by itself, an open string theory is not and requires +closed strings. +Spectrum +In order to gain some intuition for the states described by a closed string, one can write the +Fourier expansion of the fields Xµ (in the gauge gab = ηab and after imposing the equations +3We focus mainly on the bosonic string theory, leaving aside the superstring, except when differences +are important. +14 + +(a) Open string +(b) Closed string +Figure 1.3: Open and closed strings. +−−−−−−−−→ +Figure 1.4: Trajectory xµ +c (τ, σ) of a closed string in spacetime (worldsheet). It begins and +ends at the circles parametrized by xi(σ) and xf(σ). +The worldsheet is topologically a +cylinder and is parametrized by (τ, σ) ∈ [τi, τf] × [0, 2π). +−−−−−−−−→ +Figure 1.5: Trajectory xµ +c (τ, σ) of an open string in spacetime (worldsheet). +It begins +and ends at the lines parametrized by xi(σ) and xf(σ). The worldsheet is topologically a +rectangle and is parametrized by (τ, σ) ∈ [τi, τf] × [0, ℓ]. +15 + +of motion) +Xµ(τ, σ) ∼ xµ + pµτ + +i +√ +2 +� +n∈Z∗ +1 +n +� +αµ +ne−in(τ−σ) + ¯αµ +ne−in(τ+σ)� +, +(1.1) +where xµ is the centre-of-mass position of the string and pµ its momentum.4 +Canonical +quantization leads to the usual commutator: +[xµ, pν] = iηµν . +(1.2) +With respect to a point-particle for which only the first two terms are present, there are an +infinite number of oscillators αµ +n and ¯αµ +n which satisfy canonical commutation relations for +creation n < 0 and annihilation operators n > 0 +[αµ +m, αν +n] = m ηµνδm+n,0 . +(1.3) +The non-zero modes are the Fourier modes of the excitations of the embedded string. The +case of the open string is simply obtained by setting ¯αn = αn and p → 2p. The Hamiltonian +for the closed and open strings read respectively +Hclosed = −m2 +2 + N + ¯N − 2 , +(1.4a) +Hopen = −m2 + N − 1 +(1.4b) +where m2 = −pµpµ is the mass of the state (in Planck units), N and ¯N (level operators) +count the numbers Nn and ¯Nn of oscillators αn and ¯αn weighted by their mode index n: +N = +� +n∈N +nNn , +Nn = 1 +n α−n · αn , +¯N = +� +n∈N +n ¯Nn , +¯Nn = 1 +n ¯α−n · ¯αn . +(1.5) +With these elements, the Hilbert space of the string theory can be constructed. Invariance +under reparametrization leads to the on-shell condition, which says that the Hamiltonian +vanishes: +H |ψ⟩ = 0 +(1.6) +for any physical state |ψ⟩. Another constraint for the closed string is the level-matching +condition +(N − ¯N) |ψ⟩ = 0 . +(1.7) +It can be understood as fixing an origin on the string. +The ground state |k⟩ with momentum k is defined to be the eigenstate of the momentum +operator which does not contain any oscillator excitation: +pµ |k⟩ = kµ |k⟩ , +∀n > 0 : +αµ +n |k⟩ = 0 . +(1.8) +A general state can be built by applying successively creation operators +|ψ⟩ = +� +n>0 +D−1 +� +µ=0 +(αµ +−n)Nn,µ |k⟩ , +(1.9) +4In the introduction, we set α′ = 1. +16 + +where Nn,µ ∈ N counts excitation level of the oscillator αµ +−n. In the rest of this section, we +describe the first two levels of states. +The ground state is a tachyon (faster-than-light particle) because the Hamiltonian con- +straint shows that it has a negative mass (in the units where α′ = 1): +closed : +m2 = −4 , +open : +m2 = −1 . +(1.10) +The first excited state of the open string is found by applying α−1 on the vacuum |k⟩: +αµ +−1 |k⟩ . +(1.11) +This state is massless: +m2 = 0 +(1.12) +and since it transforms as a Lorentz vector (spin 1), it is identified with a U(1) gauge boson. +Writing a superposition of such states +|A⟩ = +� +dDk Aµ(k) αµ +−1 |k⟩ , +(1.13) +the coefficient Aµ(k) of the Fourier expansion is interpreted as the spacetime field for the +gauge boson. Reparametrization invariance is equivalent to the equation of motion +k2Aµ = 0 . +(1.14) +One can prove that the field obeys the Lorentz gauge condition +kµAµ = 0 , +(1.15) +which results from gauge fixing the U(1) gauge invariance +Aµ −→ Aµ + kµλ . +(1.16) +It can also be checked that the low-energy action reproduces the Maxwell action. +The first level of the closed string is obtained by applying both α−1 and ¯α−1 (this is the +only way to match N = ¯N at this level) +αµ +−1¯αν +−1 |k⟩ +(1.17) +and the corresponding states are massless +m2 = 0 . +(1.18) +These states can be decomposed into irreducible representations of the Lorentz group +� +αµ +−1¯αν +−1 + αν +−1¯αµ +−1 − 1 +D ηµνα−1 · ¯α−1 +� +|p⟩ , +� +αµ +−1¯αν +−1 − αν +−1¯αµ +−1 +� +|p⟩ , +1 +D ηµναµ +−1¯αν +−1 |p⟩ +(1.19) +which are respectively associated to the spacetime fields Gµν (metric, spin 2), Bµν (Kalb– +Ramond 2-form) and Φ (dilaton, spin 0). The appearance of a massless spin 2 particle (with +low-energy action being the Einstein–Hilbert action) is a key result and originally raised +interest for string theory. +17 + +Remark 1.1 (Reparametrization constraints) Reparametrization invariance leads to +other constraints than H = 0. They imply in particular that the massless fields have the +correct gauge invariance and hence the correct degrees of freedom. +Note that, after taking into account these constraints, the remaining modes correspond +to excitations of the string in the directions transverse to it. +Hence, each vibrational mode of the string corresponds to a spacetime field for a point- +particle (and linear superpositions of modes can describe several fields). This is how string +theory achieves unification since a single type of string (of each topology) is sufficient for +describing all the possible types of fields encountered in the standard model and in gravity. +They correspond to the lowest excitation modes, the higher massive modes being too heavy +to be observed at low energy. +Bosonic string theory includes tachyons and is thus unstable. While the instability of +the open string tachyon is well understood and indicates that open strings are unstable and +condense to closed strings, the status of the closed string tachyon is more worrisome (literally +interpreted, it indicates a decay of spacetime itself). In order to solve this problem, one can +introduce supersymmetry: in this case, the spectrum does not include the tachyon because +it cannot be paired with a supersymmetric partner. +Moreover, as its name indicates, the bosonic string possesses only bosons in its spectrum +(perturbatively), which is an important obstacle to reproduce the standard model. +By +introducing spacetime fermions, supersymmetry also solves this problem. The last direct +advantage of the superstring is that it reduces the number of dimensions from 26 to 10, +which makes the compactification easier. +1.2.2 +Classification of superstring theories +In this section, we describe the different superstring theories (Chapter 17). +In order to +proceed, we need to introduce some new elements. +The worldsheet field theory of the closed string is made of two sectors, called the left- and +right-moving sectors (the αn and ¯αn modes). While they are treated symmetrically in the +simplest models, they are in fact independent (up to the zero-mode) and the corresponding +CFT can be chosen to be distinct. +The second ingredient already evoked earlier is supersymmetry. This symmetry associ- +ates a fermion to each boson (and conversely) through the action of a supercharge Q +|boson⟩ = Q |fermion⟩ . +(1.20) +More generally, one can consider N supercharges which build up a family of several bosonic +and fermionic partners. Since each supercharge increases the spin by 1/2 (in D = 4), there +is an upper limit for the number of supersymmetries – for interacting theories with a finite +number of fields5 – in order to keep the spin of a family in the range where consistent actions +exist: +• Nmax = 4 without gravity (−1 ≤ spin ≤ 1); +• Nmax = 8 with gravity (−2 ≤ spin ≤ 2). +This counting serves as a basis to determine the maximal number of supersymmetries in +other dimensions (by relating them through dimensional reductions). +Let’s turn our attention to the case of the two-dimensional worldsheet theory. +The +number of supersymmetries of the closed left- and right-moving sectors can be chosen in- +dependently, and the number of charges is written as (NL, NR) (the index is omitted when +5These conditions exclude the cases of free theories and higher-spin theories. +18 + +statements are made at the level of the CFT). The critical dimension (absence of quantum +anomaly for the Weyl invariance) depends on the number of supersymmetry +D(N = 0) = 26, +D(N = 1) = 10. +(1.21) +Type II superstrings have (NL, NR) = (1, 1) and come in two flavours called IIA and IIB +according to the chiraly of the spacetime gravitini chiralities. A theory is called heterotic if +NL > NR; we will mostly be interested in the case NL = 1 and NR = 0.6 In such theories, +there cannot be open strings since both sectors must be equal in the latter. Since the critical +dimensions of the two sectors do not match, one needs to get rid of the additional dimensions +of the right-moving sector; this leads to the next topic – gauge groups. +Gauge groups associated with spacetime gauge bosons appear in two different places. In +heterotic models, the compactification of the unbalanced dimensions of the left sector leads +to the appearance of a gauge symmetry. The possibilities are scarce due to consistency +conditions which ensure a correct gluing with the right-sector. Another possibility is to +add degrees of freedom – known as Chan–Paton indices – at the ends of open strings: +one end transforms in the fundamental representation of a group G, while the other end +transforms in the anti-fundamental. The modes of the open string then reside in the adjoint +representation, and the massless spin-1 particles become the gauge bosons of the non-Abelian +gauge symmetry. +Finally, one can consider oriented or unoriented strings. An oriented string possesses an +internal direction, i.e. there is a distinction between going from the left to the right (for an +open string) or circling in clockwise or anti-clockwise direction (for a closed string). Such +an orientation can be attributed globally to the spacetime history of all strings (interacting +or not). The unoriented string is obtained by quotienting the theory by the Z2 worldsheet +parity symmetry which exchanges the left- and right-moving sectors. Applying this to the +type IIB gives the type I theory. +The tachyon-free superstring theories together with the bosonic string are summarized +in Table 1.1. +worldsheet +susy +D +spacetime +susy +gauge group +open string +oriented +tachyon +bosonic +(0, 0) +26 +0 +any* +yes +yes / no +yes +type I +(1, 1) +10 +(1, 0) +SO(32) +yes +no +no +type IIA +(1, 1) +10 +(1, 1) +U(1) +(yes)† +yes +no +type IIB +(1, 1) +10 +(2, 0) +none +(yes)† +yes +no +heterotic SO(32) +(1, 0) +10 +(1, 0) +SO(32) +no +yes +no +heterotic E8 +(1, 0) +10 +(1, 0) +E8 × E8 +no +yes +no +heterotic SO(16) +(1, 0) +10 +(0, 0) +SO(16) × SO(16) +no +yes +no +* UV divergences beyond the tachyon (interpreted as closed string dilaton tadpoles) cancel only for the unoriented +open plus closed strings with gauge group SO(213) = SO(8192). +† The parenthesis indicates that type II theories don’t have open strings in the vacuum: they require a D-brane +background. This is expected since there is no gauge multiplet in d = 10 (1, 1) or (2, 0) supergravities (the +D-brane breaks half of the supersymmetry). +Table 1.1: List of the consistent tachyon-free (super)string theories. The bosonic theory is +added for comparison. There are additional heterotic theories without spacetime supersym- +metry, but they contain a tachyon and are thus omitted. +6The case NL < NR is identical up to exchange of the left- and right-moving sectors. +19 + +(a) Closed strings +(b) Open strings +Figure 1.6: Graphs corresponding to 1-loop 4-point scattering after a conformal mapping. +1.2.3 +Interactions +Worldsheet and Riemann surfaces +After having described the spectrum and the general characteristics of string theory comes +the question of interactions. The worldsheets obtained in this way are Riemann surfaces, +i.e. 1-dimensional complex manifolds. They are classified by the numbers of handles (or +holes) g (called the genus) and external tubes n. In the presence of open strings, surfaces +have boundaries: in addition to the handles and tubes, they are classified by the numbers +of disks b and of strips m.7 A particularly important number associated to each surface is +the Euler characteristics +χ = 2 − 2g − b , +(1.22) +which is a topological invariant. It is remarkable that there is a single topology at every +loop order when one considers only closed strings, and just a few more in the presence of +open strings. The analysis is greatly simplified in contrast to QFT, for which the number +of Feynman graphs increases very rapidly with the number of loops and external particles. +Due to the topological equivalence between surfaces, a conformal map can be used in +order to work with simpler surfaces. In particular, the external tubes and strips are collapsed +to points called punctures (or marked points) on the corresponding surfaces or boundaries. +A general amplitude then looks like a sphere from which holes and disks have been removed +and to which marked points have been pierced (Figure 1.7). +Amplitudes +In order to compute an amplitude for the scattering of n strings (Chapters 3 and 4), one +must sum over all the inequivalent worldsheets through a path integral weighted by the CFT +action chosen to describe the theory.8 At fixed n, the sum runs over the genus g, such that +each term is described by a Riemann surface Σg,n of genus g with n punctures. +The interactions between strings follow from the graph topologies: since the latter are not +encoded into the action, the dependence in the coupling constant must be added by hand. +For closed strings, there is a unique cubic vertex with coupling gs. A direct inspection shows +that the correct factor is gn−2+2g +s +: +7We ignore unoriented strings in this discussion. +They would lead to an additional object called a +cross-cap, which is a place where the surface looses its orientation. +8For simplicity we focus on closed string amplitudes in this section. +20 + +Figure 1.7: General Riemann surfaces with boundaries and punctures. +• for n = 3 there is one factor gs, and every additional external string leads to the +addition of one vertex with factor gs, since this process can be obtained from the n−1 +process by splitting one of the external string in two by inserting a vertex; +• each loop comes with two vertices, so g-loops provide a factor g2g +s . +Remark 1.2 (Status of gs as a parameter) It was stated earlier that string theory has +no dimensionless parameter, but gs looks to be one. In reality it is determined by the expect- +ation value of the dilaton gs = e⟨Φ⟩. Hence the coupling constant is not a parameter defining +the theory but is rather determined by the dynamics of the theory. +Finally, the external states must be specified: this amounts to prescribe boundary condi- +tions for the path integral or to insert the corresponding wave functions. Under the conformal +mapping which brings the external legs to punctures located at zi, the states are mapped to +local operators Vi(ki, zi) inserted at the points zi. The latter are built from the CFT fields +and are called vertex operators: they are characterized by a momentum kµ which comes +from the Fourier transformation of the Xµ fields representing the non-compact dimensions. +These operators are inserted inside the path integral with integrals over the positions zi in +order to describe all possible conformal mappings. +Ultimately, the amplitude (amputated Green function) is computed as +An(k1, . . . , kn) = +� +g≥0 +gn−2+2g +s +Ag,n +(1.23) +where +Ag,n = +� +n +� +i=1 +d2zi +� +dgabdΨ e−Scft[gab,Ψ] +n +� +i=1 +Vi(ki, zi) +(1.24) +is the g-loop n-point amplitude (for simplicity we omit the dependence on the states beyond +the momentum). Ψ denotes collectively the CFT fields and gab is the metric on the surface. +The integration over the metrics and over the puncture locations contain a huge redund- +ancy due to the invariance under reparametrizations, which means that one integrates over +many equivalent surfaces. To avoid this, Faddeev–Popov ghosts must be introduced and the +integral is restricted to only finitely many (real) parameters tλ. They form the moduli space +Mg,n of the Riemann surfaces Σg,n whose dimension is +dimR Mg,n = 6g − 6 + 2n. +(1.25) +21 + +The computation of the amplitude Ag,n can be summarized as: +Ag,n = +� +Mg,n +6g−6+2n +� +λ=1 +dtλ F(t). +(1.26) +The function F(t) is a correlation function in the worldsheet CFT defined on the Riemann +surface Σg,n +F(t) = +� n +� +i=1 +Vi × ghosts × super-ghosts +� +Σg,n +. +(1.27) +Note that the (super)ghost part is independent of the choice of the matter CFT. +Divergences and Feynman graphs +Formally the moduli parameters are equivalent to Schwinger (proper-time) parameters si in +usual QFT: these are introduced in order to rewrite propagators as +1 +k2 + m2 = +� ∞ +0 +ds e−s(k2+m2), +(1.28) +such that the integration over the momentum k becomes a Gaussian times a polynomial. +This form of the propagator is useful to display the three types of divergences which can be +encountered: +1. IR: regions si → ∞ (for k2 + m2 ≤ 0). These divergences are artificial for k2 + m2 < 0 +and means that the parametrization is not appropriate. Divergences for k2 + m2 = 0 +are genuine and translates the fact that quantum effects shift the vacuum and the +masses. +Taking these effects into account necessitates a field theory framework in +which renormalization can be used. +2. UV: regions si → 0 (after integrating over k). Such divergences are absent in string +theories because these regions are excluded from the moduli space Mg,n (see Figure 1.8 +for the example of the torus).9 +3. Spurious: regions with finite si where the amplitude diverges. This happens typically +only in the presence of super-ghosts and it translates a breakdown of the gauge fixing +condition.10 Since these spurious singularities of the amplitudes are not physical, one +needs to ensure that they can be removed, which is indeed possible to achieve. +Hence, only IR divergences present a real challenge to string theory. Dealing with these +divergences requires renormalizing the amplitudes, but this is not possible in the standard +formulation of worldsheet string theory since the states are on-shell.11 +9There is a caveat to this statement: UV divergences reappear in string field theory in Lorentzian +signature due to the way the theory is formulated. +The solution requires a generalization of the Wick +rotation. +Moreover, this does not hold for open strings whose moduli spaces contains those regions: in this case, +the divergences are reinterpreted in terms of closed strings propagating. +10Such spurious singularities are also found in supergravity. +11The on-shell condition is a consequence of the BRST and conformal invariance. While the first will be +given up, the second will be maintained to facilitate the computations. +22 + +Figure 1.8: Moduli space of the torus: Re τ ∈ [−1/2, 1/2], Im τ > 0 and |τ| > 1. +1.3 +String field theory +1.3.1 +From the worldsheet to field theory +The first step is to solve the IR divergences problem is to go off-shell (Chapters 11 and 13). +This is made possible by introducing local coordinates around the punctures of the Riemann +surface (Chapter 12). +The IR divergences originate from Riemann surfaces close to degeneration, that is, sur- +faces with long tubes. The latter can be of separating and non-separating types, depending +on whether the Riemann surface splits in two pieces if the tube is cut (Figure 1.9). By +exploring the form of the amplitudes in this limit (Chapter 14), the expression naturally +separates into several pieces, to be interpreted as two amplitudes (of lower n and g) con- +nected by a propagator. The latter can be reinterpreted as a standard (k2 + m2)−1 term, +hence solving the divergence problem for k2 + m2 < 0. Taking this decomposition seriously +leads to identify each contribution with a Feynman graph. +Decomposing the amplitude recursively, the next step consists in finding the elementary +graphs, i.e. the interaction vertices from which all other graphs (and amplitudes) can be +built. These graphs are the building blocks of the field theory (Chapter 15), with the kinetic +term given by the inverse of the propagator. Having Feynman diagrams and a field theory +allows to use all the standard tools from QFT. +However, this field theory is gauge fixed because on-shell amplitudes are gauge invariant +and include only physical states. +For this reason, one needs to find how to re-establish +the gauge invariance. Due to the complicated structure of string theory, the full-fledged +Batalin–Vilkovisky (BV) formalism must be used (Chapter 15): it basically amounts to +introduce ghosts before the gauge fixing. The final stage is to obtain the 1PI effective action +from which the physics is more easily extracted. But, it is useful to study first the free +theory (Chapters 9 and 10) to gain some insights. The book ends with a discussion of the +momentum-space representation and of background independence (Chapters 16 and 18). +The procedure we will follow is a kind of reverse-engineering: we know what is the final +23 + +(a) Separating. +(b) Non-separating. +Figure 1.9: Degeneration of Riemann surfaces. +result and we want to study backwards how it is obtained: +on-shell amplitude → off-shell amplitude → Feynman graphs +→ gauge fixed field theory → BV field theory +In standard QFT, one follows the opposite process. +Remark 1.3 There are some prescriptions (using for example analytic continuation, the +optical theorem, some tricks. . . ) to address the problems mentioned above, but there is no +general and universally valid procedure. A field theory is much more satisfactory because it +provides a unique and complete framework. +We can now summarise the disadvantages of the worldsheet approach over the spacetime +field one: +• no natural description of (relativistic) multi-particle states; +• on-shell states: +– lack of renormalization, +– presence of infrared divergences, +– scattering amplitudes only for protected states; +• interactions added by hand; +• hard to check consistency (unitarity, causality. . . ); +• absence of non-perturbative processes. +Some of these problems can be addressed with various prescriptions, but it is desirable +to dispose of a unified and systematic procedure, which is to be found in the field theory +description. +24 + +1.3.2 +String field action +A string field theory (SFT) for open and closed strings is based on two fields Φ[X(σ)] (open +string field) and Ψ[X(σ)] (closed string field) governed by some action S[Φ, Ψ]. This action +is built from a diagonal kinetic term +S0 = 1 +2 KΨ(Ψ, Ψ) + 1 +2 KΦ(Φ, Φ) +(1.29) +and from an interaction polynomial in the fields +Sint = +� +m,n +Vm,n(Φm, Ψn) +(1.30) +where Vm,n is an appropriate product mapping m closed and n open string states to a number +(the power is with respect to the tensor product). In particular, it contains the coupling +constant. Contrary to the worldsheet approach where the cubic interaction looks sufficient, +higher-order elementary interactions with m, n ∈ N are typically needed. A second specific +feature is that the products also admit a loop (or genus g) expansion: a fundamental n- +point interaction is introduced at every loop order g. These terms are interpreted as (finite) +counter-terms needed to restore the gauge invariance of the measure. These two facts come +from the decomposition of the moduli spaces in pieces (Section 1.2.3). +Writing an action for a field Ψ[X(σ)] for which reparametrization invariance holds is +highly complicated. The most powerful method is to introduce a functional dependence in +ghost fields Ψ[X(σ), c(σ)] and to extend the BRST formalism to the string field, leading +ultimately to the BV formalism. +While the latter formalism is the most complete and +ensures that the theory is consistent at the quantum level, it is difficult to characterize the +interactions explicitly. Several constructions which exploit different properties of the theory +have been proposed: +• direct computation by reverse engineering of worldsheet amplitudes; +• specific parametrization of the Riemann surfaces (hyperbolic, minimal area); +• analogy with Chern–Simons and Wess–Zumino–Witten (WZW) theories; +• exploitation of the L∞ and A∞ algebra structures. +It can be shown that these constructions are all equivalent. For the superstring, the simplest +strategy is to dress the bosonic interactions with data from the super-ghost sector, which +motivates the study of the bosonic SFT by itself. The main difficulty in working with SFT is +that only the first few interactions have been constructed explicitly. Finally, the advantage +of the first formulation is that it provides a general formulation of SFT at the quantum +level, from which the general structure can be studied. +1.3.3 +Expression with spacetime fields +To obtain a more intuitive picture and to make contact with the spacetime fields, the field +is expanded in terms of 1-particle states in the momentum representation +|Ψ⟩ = +� +n +� +dDk +(2π)D ψα(k) |k, α⟩ , +(1.31) +where α denotes collectively the discrete labels of the CFT eigenstates. The coefficients +ψα(k) of the CFT states |k, α⟩ are spacetime fields, the first ones being the same as those +found in the first-quantized picture (Section 1.2.1) +ψα = {T, Gµν, Bµν, Φ, . . .}. +(1.32) +25 + +Then, inserting this expansion in the action gives an expression like S[T, Gµν, . . .]. The exact +expression of this action is out of reach and only the lowest terms are explicitly computable +for a given CFT background. Nonetheless, examining the string field action indicates what +is the generic form of the action in terms of the spacetime fields. One can then study the +properties of such a general QFT: since it is more general than the SFT (expanded) action, +any result derived for it will also be valid for SFT. This approach is very fruitful for studying +properties related to consistency of QFT (unitarity, soft theorems. . . ) and this can provide +helpful phenomenological models. +In conclusion, SFT can be seen as a regular QFT with the following properties: +• infinite number of fields; +• non-local interaction (proportional to e−k2#); +• the amplitudes agree with the worldsheet amplitudes (when the latter can be defined); +• genuine (IR) divergences agree but can be handled with the usual QFT tools. +1.3.4 +Applications +The first aspect is the possibility to use standard QFT techniques (such as renormalization) +to study – and to make sense of – string amplitudes. In this sense, SFT can be viewed as +providing recipes for computing quantities in the worldsheet theory which are otherwise not +defined. This program has been pushed quite far in the last years. +Another reason to use SFT is gauge invariance: it is always easier to describe a sys- +tem when its gauge invariance is manifest. We have explained that string theory contains +Yang–Mills and graviton fields with the corresponding (spacetime) gauge invariances (non- +Abelian gauge symmetry and diffeomorphisms). In fact, these symmetries are enhanced to +an enormous gauge invariance when taking into account the higher-spin fields. This invari- +ance is hidden in the standard formulation and cannot be exploited fully. On the other +hand, the full gauge symmetry is manifest in string field theory. +Finally, the worldvolume description of p-brane is difficult because there is no analogue +of the Polyakov action. If one could find a first-principle description of SFT which does not +rely on CFT and first-quantization, then one may hope to generalize it to build a brane field +theory. +We can summarize the general motivations for studying SFT: +• field theory (second-quantization); +• more rigorous and constructive formulation; +• make gauge invariance explicit (L∞ algebras et al.); +• use standard QFT techniques (renormalization, analyticity. . . ) +→ remove IR divergences, prove consistency (Cutkosky rules, unitarity, soft theorems, +background independence. . . ); +• worldvolume theory ill-defined for (p > 1)-branes. +Beyond these general ideas, SFT has been developed in order to address different questions: +• worldsheet scattering amplitudes; +• effective actions; +• map of the consistent backgrounds (classical solutions, marginal deformations, RR +fluxes. . . ); +26 + +• collective, non-perturbative, thermal, dynamical effects; +• symmetry breaking effects; +• dynamics of compactification; +• proof of dualities; +• proof of the AdS/CFT correspondence. +The last series of points is still out of reach within the current formulation of SFT. However, +the last two decades have seen many important develoments developments: +• construction of the open, closed and open-closed superstring field theories: +– 1PI and BV actions and general properties [73, 165, 166, 213, 214, 216, 218, 220, +222, 225, 226, 230], +– dressing of bosonic products using the WZW construction and homotopy al- +gebra [18, 19, 67–70, 74–76, 79, 80, 94, 111, 131, 134, 140–146, 180], +– light-cone super-SFT [114–117], +– supermoduli space [175, 241]; +• hyperbolic and minimal area constructions [38, 102, 103, 162–164, 183]; +• open string analytic solutions [77, 78]; +• level-truncation solutions [135–137]; +• field theory properties [34, 43, 150, 187, 221, 223, 224]; +• spacetime effective actions [65, 153, 154, 248]; +• defining worldsheet scattering amplitudes [184–186, 215, 216, 219, 227–229]; +• marginal and RR deformations [35, 229, 248]. +Recent reviews are [42, 71, 72]. +1.4 +Suggested readings +For references about different aspects in this chapter: +• Differences between the worldvolume and spacetime formalisms – and of the associated +first- and second-quantization – for the particle and string [124, chap. 1, 265, chap. 11]. +• General properties of relativistic strings [92, 265]. +• Divergences in string theory [42, 217, 256, sec. 7.2]. +• Motivations for building a string field theory [192, sec. 4]. +27 + +Part I +Worldsheet theory +28 + +Chapter 2 +Worldsheet path integral: +vacuum amplitudes +Abstract +In this chapter, we develop the path integral quantization for a generic closed +string theory in worldsheet Euclidean signature. We focus on the vacuum amplitudes, leaving +scattering amplitudes for the next chapter. This allows to focus on the definition and gauge +fixing of the path integral measure. +The exposition differs from most traditional textbooks in three ways: 1) we consider a +general matter CFT, 2) we consider the most general treatment (for any genus) and 3) we +don’t use complex coordinates but always a covariant parametrization. +The derivation is technical and the reader is encouraged to not stop at this chapter +in case of difficulties and to proceed forward: most concepts will be reintroduced from a +different point of view later in other chapters of the book. +2.1 +Worldsheet action and symmetries +The string worldsheet is a Riemann surface W = Σg of genus g: the genus counts the +number of holes or handles. Coordinates on the worldsheet are denoted by σa = (τ, σ). +When there is no risk of confusion, σ denotes collectively both coordinates. Since closed +strings are considered, the Riemann surface has locally the topology of a cylinder, with the +spatial section being circles S1 with radius taken to be 1, such that +σ ∈ [0, 2π), +σ ∼ σ + 2π. +(2.1) +The string is embedded in the D-dimensional spacetime M with metric Gµν through maps +Xµ(σa) : W → M with µ = 0, . . . , D − 1. +The Nambu–Goto action is the starting point of the worldsheet description: +SNG[Xµ] = +1 +2πα′ +� +d2σ +� +det Gµν(X)∂Xµ +∂σa +∂Xν +∂σb , +(2.2) +where α′ is the Regge slope (related to the string tension and string length). +However, +quantizing this action is difficult because it is highly non-linear. To solve this problem, a +Lagrange multiplier is introduced to remove the squareroot. This auxiliary field corresponds +to an intrinsic worldsheet metric gab(σ). +The worldsheet dynamics is described by the +Polyakov action: +SP[g, Xµ] = +1 +4πα′ +� +d2σ√g gabGµν(X)∂Xµ +∂σa +∂Xν +∂σb , +(2.3) +29 + +which is classically equivalent to the Nambu–Goto action (2.2). In this form, it is clear that +the scalar fields Xµ(σ) (µ = 0, . . . D − 1) characterize the string theory under consideration +in two ways. +First, by specifying some properties of the spacetime in which the string +propagates (for example, the number of dimensions is determined by the number of fields +Xµ), second, by describing the internal degrees of freedom (vibration modes).1 +But, nothing prevents to consider a more general matter content in order to describe a +different spacetime or different degrees of freedom. In Polyakov’s formalism, the worldsheet +geometry is endowed with a metric gab(σ) together with a set of matter fields living on it. +The scalar fields Xµ can be described by a general sigma model which encodes the embedding +of the string in the D non-compact spacetime dimensions, and other fields can be added, +for example to describe compactified dimensions or (spacetime) spin. Different sets of fields +(and actions) correspond to different string theories. +However, to describe precisely the +different possibilities, we first have to understand the constraints on the worldsheet theories +and to introduce conformal field theories (Part I). In this chapter (and in most of the book), +the precise matter content is not important and we will denote the fields collectively as Ψ(σ). +Before discussing the symmetries, let’s introduce a topological invariant which will be +needed throughout the text: the Euler characteristics. It is computed by integrating the +Riemann curvature R of the metric gab over the surface Σg: +χg := χ(Σg) := 2 − 2g = 1 +4π +� +Σg +d2σ√g R, +(2.4) +where g is the genus of the surface. Oriented Riemann surfaces without boundaries are +completely classified (topologically or as complex manifolds) by their Euler characteristics +χg, or equivalently by their genus g. +In order to describe a proper string theory, the worldsheet metric gab(σ) should not +be dynamical. +This means that the worldsheet has no intrinsic dynamics and that no +supplementary degrees of freedom are introduced when parametrizing the worldsheet with +a metric. A solution to remove these degrees of freedom is to introduce gauge symmetries +with as many gauge parameters as there are of degrees of freedom. The simplest symmetry +is invariance under diffeomorphisms: indeed, the worldsheet theory is effectively a QFT +coupled to gravity and it makes sense to require this invariance. Physically, this corresponds +to the fact that the worldsheet spatial coordinate σ used along the string and worldsheet +time are arbitrary. However, diffeomorphisms alone are not sufficient to completely fix the +metric. Another natural candidate is Weyl invariance (local rescalings of the metric). +A diffeomorphism f ∈ Diff(Σg) acts on the fields as +σ′a = f a(σb), +g′(σ′) = f ∗g(σ), +Ψ′(σ′) = f ∗Ψ(σ), +(2.5) +where the star denotes the pullback by f: this corresponds simply to the standard coordinate +transformation where each tensor index of the field receives a factor ∂σa/∂σ′b. In particular, +the metric and scalar fields transform explicitly as +g′ +ab(σ′) = ∂σc +∂σ′a +∂σd +∂σ′b gcd(σ), +X′µ(σ′) = Xµ(σ). +(2.6) +The index µ is inert since it is a target spacetime index: from the worldsheet point of view, +it just labels a collection of worldsheet scalar fields. Infinitesimal variations are generated +by vector fields on Σg: +δξσa = ξa, +δξΨ = LξΨ, +δξgab = Lξgab, +(2.7) +1Obviously, the vibrational modes are also constrained by the spacetime geometry. +30 + +where Lξ is the Lie derivative2 with respect to the vector field ξ ∈ diff(Σg) ≃ TΣg. The Lie +derivative of the metric is +Lξgab = ξc∂cgab + gac∂bξc + gbc∂aξc = ∇aξb + ∇bξa. +(2.8) +The Lie algebra generates only transformations in the connected component Diff0(Σg) of +the diffeomorphism group which contains the identity. +Transformations not contained in Diff0(Σg) are called large diffeomorphisms: this in- +cludes reflections, for example. The quotient of the two groups is called the modular group +Γg (also mapping class group or MCG): +Γg := π0 +� +Diff(Σg) +� += Diff(Σg) +Diff0(Σg). +(2.9) +It depends only on the genus g of the Riemann surface, but not on the metric. It is an +infinite discrete group for genus g ≥ 1 surfaces; in particular, Γ1 = SL(2, Z). +A Weyl transformation e2ω ∈ Weyl(Σg) corresponds to a local rescaling of the metric +and leaves the other fields unaffected3 +g′ +ab(σ) = e2ω(σ)gab(σ), +Ψ′(σ) = Ψ(σ). +(2.10) +The exponential parametrization is generally more useful, but one should remember that it +is e2ω and not ω which is an element of the group. The infinitesimal variation reads +δωgab = 2ω gab, +δωΨ = 0 +(2.11) +where ω ∈ weyl(Σ) ≃ F(Σg) is a function on the manifold. Two metrics related in this way +are said to be conformally equivalent. The conformal structure of the Riemann surface is +defined by +Conf(Σg) := Met(Σg) +Weyl(Σg), +(2.12) +where Met(Σg) denotes the space of all metrics on Σg. Each element is a class of conformally +equivalent metrics. +Diffeomorphisms have two parameters ξa (vector field) and Weyl invariance has one, +ω (function). +Hence, this is sufficient to locally fix the three components of the metric +(symmetric matrix) and the total gauge group of the theory is the semi-direct product +G := Diff(Σg) ⋉ Weyl(Σg). +(2.13) +Similarly, the component connected of the identity is written as +G0 := Diff0(Σg) ⋉ Weyl(Σg). +(2.14) +The semi-direct product arises because the Weyl parameter is not inert under diffeo- +morphisms. Indeed, the combination of two transformations is +g′ = f ∗� +e2ωg +� += e2f ∗ωf ∗g, +(2.15) +such that the diffeomorphism acts also on the conformal factor. +2For our purpose here, it is sufficient to accept the definition of the Lie derivative as corresponding to +the infinitesimal variation. +3For simplicity, we consider only fields which do not transform under Weyl transformations, which +excludes fermions. +31 + +The combination of transformations (2.15) can be chosen to fix the metric in a convenient +gauge. For example, the conformal gauge reads +gab(σ) = e2φ(σ)ˆgab(σ), +(2.16) +where ˆgab is some (fixed) background metric and φ(σ) is the conformal factor, also called +the Liouville field. Fixing only diffeomorphisms amount to keep φ arbitrary: the latter can +then be fixed with a Weyl transformation. For instance, one can adopt the conformally flat +gauge +ˆgab = δab, +φ arbitrary +(2.17) +with a diffeomorphism, and then reach the flat gauge +ˆgab = δab, +φ = 0 +(2.18) +with a Weyl transformation. Another common choice is the uniformization gauge where ˆg +is taken to be the metric of constant curvature on the sphere (g = 0), on the plane (g = 1) +or on the hyperbolic space (g > 1). All these gauges are covariant (both in spacetime and +worldsheet). +Remark 2.1 (Active and passive transformations) Usually, symmetries are described +by active transformations, which means that the field is seen to be changed by the transform- +ation. On the other hand, gauge fixing is seen as a passive transformation, where the field +is expressed in terms of other fields (i.e. a different parametrization). These are mathem- +atically equivalent since both cases correspond to inverse elements, and one can choose the +most convenient representation. We will use indifferently the same name for the parameters +to avoid introducing minus signs and inverse. +Remark 2.2 (Topology and gauge choices) While it is always possible to adopt locally +the flat gauge (2.18), it may not be possible to extend it globally. The can be seen intuitively +from the fact that the sign of the curvature is given by the one of 1 − g, but the curvature of +the flat metric is zero: curvature must then be localized somewhere and this prevents from +using a single coordinate patch. +The final step is to write an action Sm[g, Ψ] for the matter fields. According to the +previous discussion, it must have the following properties: +• local in the fields; +• renormalizable; +• non-linear sigma models for the scalar fields; +• periodicity conditions; +• invariant under diffeomorphisms (2.5); +• invariant under Weyl transformations (2.10). +The latter two conditions are summarized by +Sm[f ∗g, f ∗Ψ] = Sm[g, Ψ], +Sm[e2ωg, Ψ] = Sm[g, Ψ]. +(2.19) +The invariance under diffeomorphisms is straightforward to enforce by using only covariant +objects. Since the scalar fields represent embedding of the string in spacetime, the non- +linear sigma model condition means that spacetime is identified with the target space of the +sigma model, of which D dimensions are non-compact, and the spacetime metric appears +32 + +in the matter action as in (2.3). +The isometries of the target manifold metric become +global symmetries of Sm: while they are not needed in this chapter, they will have their +importances in other chapters. Finally, to make the action consistent with the topology of +the worldsheet, the fields must satisfy appropriate boundary conditions. For example, the +scalar fields Xµ must be periodic for the closed string: +Xµ(τ, σ) ∼ Xµ(τ, σ + 2π). +(2.20) +Remark 2.3 (2d gravity) The setup in two-dimensional gravity is exactly similar, except +that the system is, in general, not invariant under Weyl transformations. As a consequence, +one component of the metric (usually taken to be the Liouville mode) remains unconstrained: +in the conformal gauge, (2.16) only ˆg is fixed. +The symmetries (2.19) of the action have an important consequence: they imply that the +matter action is conformally invariant on flat space gab = δab. A two-dimensional conformal +field theory (CFT) is characterized by a central charge cm: roughly, it is a measure of the +quantum degrees of freedom. The central charge is additive for decoupled sectors. In partic- +ular, the scalar fields Xµ contribute as D, and it is useful to define the perpendicular CFT +with central charge c⊥ as the matter which does not describe the non-compact dimensions: +cm = D + c⊥. +(2.21) +This will be discussed in length in Part I. For this chapter and most of the book, it is +sufficient to know that the matter is a CFT of central charge cm and includes D scalar fields +Xµ: +matter CFT parameters: D, cm. +(2.22) +The energy–momentum is defined by +Tm,ab := − 4π +√g +δSm +δgab . +(2.23) +The variation of the action under the transformations (2.7) vanishes on-shell if the energy– +momentum tensor is conserved +∇aTm,ab = 0 +(on-shell). +(2.24) +On the other hand, the variation under (2.11) vanishes off-shell (i.e. without using the +equations of motion) if the energy–momentum tensor is traceless: +gabTm,ab = 0 +(off-shell). +(2.25) +The conserved charges associated to the energy–momentum tensor generate worldsheet +translations +P a := +� +dσ T 0a +m . +(2.26) +The first component is identified with the worldsheet Hamiltonian P 0 = H and generates +time translations, the second component generates spatial translations. +Remark 2.4 (Tracelessness of the energy–momentum tensor) In fact, the trace can +also be proportional to the curvature +gabTm,ab ∝ R. +(2.27) +Then, the equations of motion are invariant since the integral of R is topological. The theory +is invariant even if the action is not. Importantly, this happens for fields at the quantum +level (Weyl anomaly), for the Weyl ghost field (Section 2.4) and for the Liouville theory +(two-dimensional gravity coupled to conformal matter). +33 + +2.2 +Path integral +The quantization of the system is achieved by considering the path integral, which yields +the genus-g vacuum amplitude (or partition function): +Zg := +� +dggab +Ωgauge[g] Zm[g], +Zm[g] := +� +dgΨ e−Sm[g,Ψ] +(2.28) +at fixed genus g (not to be confused with the metric). The integration over gab is performed +over all metrics of the genus-g Riemann surface Σg: gab ∈ Met(Σg). The factor Ωgauge[g] +is a normalization inserted in order to make the integral finite: it depends on the metric +(but only through the moduli parameters, as we will show later) [53, p. 931], which explains +why it is included after the integral sign. Its value will be determined in the next section by +requiring the cancellation of the infinities due to the integration over the gauge parameters. +This partition function corresponds to the g-loop vacuum amplitude: interactions and their +associated scattering amplitudes are discussed in Section 3.1. +In order to perform the gauge fixing and to manipulate the path integral (2.28), it is +necessary to define the integration measure over the fields. Because the space is infinite- +dimensional, this is a difficult task. +One possibility is to define the measure implicitly +through Gaussian integration over the field tangent space (see also Appendix C.1). A Gaus- +sian integral involves a quadratic form, that is, an inner-product (or equivalently a metric) +on the field space. The explanation is that a metric also defines a volume form, and thus +a measure. To reduce the freedom in the definition of the inner-product, it is useful to +introduce three natural assumptions: +1. ultralocality: the measure is invariant under reparametrizations and defined point-wise, +which implies that it can depend on the fields but not on their derivatives; +2. invariant measure: the measure for the matter transforms trivially under any sym- +metry of the matter theory by contracting indices with appropriate tensors; +3. free-field measure: for fields other than the worldsheet metric and matter (like ghosts, +Killing vectors, etc.), the measure is the one of a free field. +This means that the inner-product is obtained by contracting the worldsheet indices of the +fields with a tensor built only from the worldsheet metric, by contracting other indices (like +spacetime) with some invariant tensor (like the spacetime metric), and finally by integrating +over the worldsheet. +We need to distinguish the matter fields from those appearing in the gauge fixing proced- +ure. The matter fields live in the representation of some group under which the inner-product +is invariant: this means that it is not possible to define each field measure independently +if the exponential of inner-products does not factorize. As an example, on a curved back- +ground: dX ̸= � +µ dXµ. However, we will not need to write explicitly the partition function +for performing the gauge fixing: it is sufficient to know that the matter is a CFT. In the +gauge fixing procedure, different types of fields (including the metric) appear which don’t +carry indices (beyond the worldsheet indices). Below, we focus on defining a measure for +each of those single fields (and use free-field measures according to the third condition). +Considering the finite elements δΦ1 and δΦ2 of tangent space at the point Φ of the state +of fields, the inner-product (·, ·)g and its associated norm | · |g read +(δΦ1, δΦ2)g := +� +d2σ√g γg(δΦ1, δΦ2), +|δΦ|2 +g := (δΦ, δΦ)g, +(2.29) +where γg is a metric on the δΦ space. It is taken to be flat for all fields except the metric itself, +that is, independent of Φ. The dependence in the metric ensures that the inner-product is +34 + +diffeomorphism invariant, which in turns will lead to a metric-dependent but diffeomorphism +invariant measure. The functional measure is then normalized by a Gaussian integral: +� +dgδΦ e− 1 +2 (δΦ,δΦ)g = +1 +� +det γg +. +(2.30) +This, in turn, induces a measure on the field space itself: +� +dΦ +� +det γg +(2.31) +The determinant can be absorbed in the measure, such that +� +dgδΦ e− 1 +2 (δΦ,δΦ)g = 1. +(2.32) +In fact, this normalization and the definition of the inner-product is ambiguous, but the +ultralocality condition allows to fix uniquely the final result (Section 2.3.4). Moreover, such +a free-field measure is invariant under field translations +Φ(σ) −→ Φ′(σ) = Φ(σ) + ε(σ). +(2.33) +The most natural inner-products for single scalar, vector and symmetric tensor fields are +(δf, δf)g := +� +d2σ√g δf 2 +(2.34a) +(δV a, δV a)g := +� +d2σ√g gabδV aδV b, +(2.34b) +(δTab, δTab)g := +� +d2σ√g GabcdδTabδTcd, +(2.34c) +where the (DeWitt) metric for the symmetric tensor is +Gabcd := Gabcd +⊥ ++ u gabgcd, +Gabcd +⊥ +:= gacgbd + gadgbc − gabgcd, +(2.35) +with u a constant. The first term G⊥ is the projector on the traceless component of the +tensor. Indeed, consider a traceless tensor gabTab = 0 and a pure trace tensor Λgab, then we +have: +GabcdTcd = Gabcd +⊥ +Tcd = 2Tab, +Gabcd(Λgcd) = 2u (Λgab). +(2.36) +While all measures are invariant under diffeomorphisms, only the vector measure is +invariant under Weyl transformations. This implies the existence of a quantum anomaly +(the Weyl or conformal anomaly): the classical symmetry is broken by quantum effects +because the path integral measure cannot respect all the classical symmetries. Hence, one +can expect difficulties for imposing it at the quantum level and ensuring that the Liouville +mode in (2.16) remains without dynamics. +The metric variation (symmetric tensor) is decomposed in its trace and traceless parts +δgab = gab δΛ + δg⊥ +ab, +δΛ = 1 +2 gabδgab, +gabδg⊥ +ab = 0. +(2.37) +In this decomposition, both terms are decoupled in the inner-product +|δgab|2 +g = 4u|δΛ|2 +g + |δg⊥ +µν| +2 +g, +(2.38) +35 + +where the norm of δΛ is the one of a scalar field (2.34a). The norm for δg⊥ +ab is equivalent +to (2.34c) with u = 0 (since it is traceless). Requiring positivity of the inner-product for a +non-traceless tensor imposes the following constraint on u: +u > 0. +(2.39) +One can absorb the coefficient with u in δΛ, which will just contribute as an overall factor: +its precise value has no physical meaning. The simple choice u = 1/4 sets the coefficient of +|δΛ|2 +g to 1 in (2.38) (another common choice is u = 1/2). Ultimately, this implies that the +measure factorizes as +dggab = dgΛ dgg⊥ +ab. +(2.40) +Computation – Equation (2.38) +Gabcd δgabδgcd = +� +Gabcd +⊥ ++ u gabgcd�� +gab δΛ + δg⊥ +ab +�� +gcd δΛ + δg⊥ +cd +� += +� +2u gcd δΛ + Gabcd +⊥ +δg⊥ +ab +�� +gcd δΛ + δg⊥ +cd +� += 4u (δΛ)2 + Gabcd +⊥ +δg⊥ +abδg⊥ +cd += 4u δΛ2 + 2gacgbdδg⊥ +abδg⊥ +cd. +Remark 2.5 Another common parametrization is +Gabcd = gacgbd + c gabgcd. +(2.41) +It corresponds to (2.35) up to a factor 1/2 and setting u = 1 + 2c. +Remark 2.6 (Matter and curved background measures) As explained previously, mat- +ter fields carry a representation and the inner-product must yield an invariant combination. +In particular, spacetime indices must be contracted with the spacetime metric Gµν(X) (which +is the non-linear sigma model metric appearing in front of the kinetic term) for a general +curved background. For example, the inner-product for the scalar fields Xµ is +(δXµ, δXµ)g = +� +d2σ√g Gµν(X)δXµδXν. +(2.42) +It is not possible to normalize anymore the measure to set det G(X) = 1 like in (2.32) since +it depends on the fields. On the other hand, this factor is not important for the manipu- +lations performed in this chapter. Any ambiguity in the measure will again corresponds to +a renormalization of the cosmological constant [53, p. 923]. Moreover, as explained above, +it is not necessary to write explicitly the matter partition function as long as it describes a +CFT. +2.3 +Faddeev–Popov gauge fixing +The naive integration over the space Met(Σg) of all metrics of Σg (note that the genus is +fixed) leads to a divergence of the functional integral since equivalent configurations +(f ∗g, f ∗Ψ) ∼ (g, Ψ), +(e2ωg, Ψ) ∼ (g, Ψ) +(2.43) +gives the same contribution to the integral. This infinite redundancy causes the integral +to diverge, and since the multiple counting is generated by the gauge group, the infinite +contribution corresponds to the volume of the latter. The Faddeev–Popov procedure is a +36 + +means to extract this volume by separating the integration over the gauge and physical +degrees of freedom +d(fields) = Jacobian × d(gauge) × d(physical). +(2.44) +The space of fields (g, Ψ) is divided into equivalence classes and one integrates over only one +representative of each class (gauge slice), see Figure 2.1. This change of variables introduces +a Jacobian which can be represented by a partition function with ghost fields (fields with +a wrong statistics). This program encounters some complications since G is a semi-direct +product and is non-connected. +Example 2.1 – Gauge redundancy +A finite-dimensional integral which mimics the problem is +Z = +� +R2 dx dy e−(x−y)2. +(2.45) +One can perform the change of variables +r = x − y, +y = a +(2.46) +such that +Z = +� +R +da +� ∞ +0 +e−r2 = +√π +2 Vol(R), +(2.47) +and Vol(R) is to be interpreted as the volume of the gauge group (translation by a real +number a). +Remark 2.7 Mathematically, the Faddeev–Popov procedure consists in identifying the or- +bits (class of equivalent metrics) under the gauge group G and to write the integral in terms +of G-invariant objects (orbits instead of individual metrics). This can be done by decompos- +ing the tangent space into variations generated by G and its complement. Then, one can +define a foliation of the field space which equips it with a fibre bundle structure: the base +is the push-forward of the complement and the fibre corresponds to the gauge orbits. The +integral is then defined by selecting a section of this bundle. +2.3.1 +Metrics on Riemann surfaces +According to the above procedure, each metric gab ∈ Met(Σg) has to be expressed in terms +of gauge parameters (ξ and ω) and of a metric ˆgab which contains the remaining gauge- +independent degrees of freedom. As there are as many gauge parameters as metric com- +ponents (Section 2.1), one could expect that there are no remaining physical parameters +and then that ˆg is totally fixed. But, this is not the case and the metric ˆg depends on a +finite number of parameters ti (moduli). The reason for this is topological: while locally it +is always possible to completely fix the metric, topological obstructions may prevent doing +it globally. This means that not all conformal classes in (2.12) can be (globally) related by +a diffeomorphism. +The quotient of the space of metrics by gauge transformations is called the moduli space +Mg := Met(Σg) +G +. +(2.48) +Accordingly, its coordinates ti with i = 1, . . . , dimR Mg are called moduli parameters. The +Teichmüller space Tg is obtained by taking the quotient of Met(Σg) with the component +37 + +Figure 2.1: The space of metrics decomposed in gauge orbits. Two metrics related by a +gauge transformation lie on the same orbit. Choosing a gauge slice amounts to pick one +metric in each orbit, and the projection gives the space of metric classes. +connected to the identity +Tg := Met(Σg) +G0 +. +(2.49) +The space Tg is the covering space of Mg: +Mg = Tg +Γg +, +(2.50) +where Γg is the modular group defined in (2.9). Both spaces can be endowed with a complex +structure and are finite-dimensional [172]: +Mg := dimR Mg = dimR Tg = +� +� +� +� +� +0 +g = 0, +2 +g = 1, +6g − 6 +g ≥ 2, +(2.51) +In particular, their volumes are related by +� +Mg +dMgt = +1 +ΩΓg +� +Tg +dMgt +(2.52) +where ΩΓg is the volume of Γg. +We will need to extract volumes of different groups, so it is useful to explain how they +are defined. A natural measure on a connected group G is the Haar measure dg, which is +the unique left-invariant measure on G. Integrating the measure gives the volume of the +group +ΩG := +� +G +dg = +� +G +d(hg), +(2.53) +for any h ∈ G. Given the Lie algebra g of the group, a general element of the algebra is a +linear combinations of the generators Ti with coefficients αi +α = αiTi. +(2.54) +38 + +Group elements can be parametrized in terms of α through the exponential map. Moreover, +since a Lie group is a manifold, it is locally isomorphic to Rn: this motivates the use of a +flat metric for the Lie algebra, such that +ΩG = +� +dα := +� � +i +dαi. +(2.55) +Finally, it is possible to perform a change of coordinates from the Lie parameters to co- +ordinates x on the group: the resulting Jacobian is the Haar measure for the coordinates +x. +Remark 2.8 While Tg is a manifold, this is not the case of Mg for g ≥ 2, which is an +orbifold: the quotient by the modular group introduces singularities [173]. +Remark 2.9 (Moduli space and fundamental domain) Given a group acting on a space, +a fundamental domain for a group is a subspace such that the full space is generated by act- +ing with the group on the fundamental domain. Hence, one can view the moduli space Mg +as a fundamental domain (sometimes denoted by Fg) for the group Γg and the space Tg. +In the conformal gauge (2.16), the metric gab can be parametrized by +gab = ˆg(f,φ) +ab +(t) := e2f ∗φf ∗ˆgab(t) = f ∗� +e2φˆgab(t) +� +(2.56) +where φ := ω and t denotes the dependence in the moduli parameters. To avoid surcharging +the notations, we will continue to write g when there is no ambiguity. In coordinates, this +is equivalent to: +gab(σ) = ˆg(f,φ) +ab +(σ; t) := e2φ(σ)ˆg′ +ab(σ; t), +ˆg′ +ab(σ; t) = ∂σ′c +∂σa +∂σ′d +∂σb ˆgcd(σ′; t). +(2.57) +Remark 2.10 Strictly speaking, the matter fields also transform and one should write Ψ = +Ψ(f) := f ∗ ˆΨ and include them in the change of integration measures of the following sections. +But, this does not bring any particular benefits since these changes are trivial because the +matter is decoupled from the metric. +Remark 2.11 Although the metric cannot be completely gauge fixed, having just a finite- +dimensional integral is much simpler than a functional integral. In higher dimensions, the +gauge fixing does not reduce that much the degrees of freedom and a functional integral over +ˆg remains (in similarity with Yang–Mills theories). +The corresponding infinitesimal transformations are parametrized by (φ, ξ, δti). +The +variation of the metric (2.56) can be expressed as +δgab = 2φ gab + ∇aξb + ∇bξa + δti∂igab, +(2.58) +which is decomposed in a reparametrization (2.7), a Weyl rescaling (2.11), and a contri- +bution from the variations of the moduli parameters. +The latter are called Teichmüller +deformations and describe changes in the metric which cannot be written as a combination +of diffeomorphism and Weyl transformation. Only the last term is written with a delta +because the parameters ξ and φ are already infinitesimal. There is an implicit sum over i +and we have defined +∂i := ∂ +∂ti +. +(2.59) +39 + +According to the formula (2.55), the volumes ΩDiff0[g] and ΩWeyl[g] of the diffeomorph- +isms connected to the identity and Weyl group are +ΩDiff0[g] := +� +dgξ, +(2.60a) +ΩWeyl[g] := +� +dgφ. +(2.60b) +The full diffeomorphism group has one connected component for each element of the modular +group Γg, according to (2.9): the volume ΩDiff[g] of the full group is the volume of the +component connected to the identity times the volume ΩΓg +ΩDiff[g] = ΩDiff0[g] ΩΓg. +(2.60c) +We have written that the volume depends on g: but, the metric itself is parametrized +in terms of the integration variables, and thus the LHS of (2.60) cannot depend on the +variable which is integrated over: ΩDiff0 can depend only on φ and ΩWeyl only on ξ. But, all +measures (2.34b) are invariant under diffeomorphisms, and thus the result cannot depend on +ξ. Moreover, the measure for vector is invariant under Weyl transformation, which means +that ΩDiff0 does not depend on φ. This implies that the volumes depend only on the moduli +parameters +ΩDiff0[g] := ΩDiff0[e2φˆg] = ΩDiff0[ˆg], +ΩWeyl[g] := ΩWeyl[Lξˆg] = ΩWeyl[ˆg]. +(2.61a) +For this reason, it is also sufficient to take the normalization factor Ωgauge to have the same +dependence: +Ωgauge[g] := Ωgauge[ˆg]. +(2.61b) +These volumes are also discussed in Section 2.3.4. +Computation – Equation (2.61) +ΩDiff0[e2φˆg] = +� +de2φLξˆgξ = +� +de2φˆgξ = +� +dˆgξ = ΩDiff0[ˆg], +ΩWeyl[Lξˆg] = +� +de2φLξˆgφ = +� +de2φˆgφ = ΩWeyl[ˆg]. +Remark 2.12 (Free-field measure for the Liouville mode) The explicit measure (2.60b) +of the Liouville mode is complicated since the inner-product contains an exponential of the +field: +|δφ|2 = +� +d2σ√g δφ2 = +� +d2σ +� +ˆg e2φδφ2. +(2.62) +It has been proposed by David–Distler–Kawai [40, 55], and later checked explicitly [50, 51, +160], how to rewrite the measure in terms of a free measure weighted by an effective action. +The latter is identified with the Liouville action (Section 2.3.3). +In principle, we could follow the standard Faddeev–Popov procedure by inserting a delta +function for the gauge fixing condition +Fab := gab − ˆg(f,φ) +ab +(t), +(2.63) +with ˆg(f,φ) +ab +(t) defined in (2.56). However, we will take a detour to take the opportunity to +study in details manipulations of path integrals and to understand several aspects of the +40 + +geometry of Riemann surfaces. In any case, several points are necessary even when going +the short way, but less apparent. +In order to make use of the factorization (2.40) of the integration measure, the variation +(2.58) is decomposed into its trace (first term) and traceless parts (last two terms) (2.37) +δgab = 2˜Λ gab + (P1ξ)ab + δti µiab, +(2.64) +where4 +(P1ξ)ab = ∇aξb + ∇bξa − gab∇cξc, +(2.65a) +µiab = ∂igab − 1 +2 gab gcd∂igcd, +(2.65b) +˜Λ = Λ + 1 +2 δti gab∂igab, +Λ = φ + 1 +2 ∇cξc. +(2.65c) +The objects µi are called Beltrami differentials and correspond to traceless Teichmüller +deformations (the factor of 1/2 comes from the symmetrization of the metric indices). The +decomposition emphasizes which variations are independent from each other. In particular, +changes to the trace of the metric due to a diffeomorphism generated by ξ or a modification +of the moduli parameters can be compensated by a Weyl rescaling. +One can use (2.40) to replace the integration over gab by one over the gauge parameters +ξ and φ and over the moduli ti since they contain all the information about the metric: +Zg = +� +dMgt dg ˜Λ dg(P1ξ) Ωgauge[g]−1 Zm[g]. +(2.66) +It is tempting to perform the change of variables +(P1ξ, ˜Λ) −→ (ξ, φ) +(2.67) +such that +dg(P1ξ) dg ˜Λ +?= dgξ dgφ ∆FP[g] +(2.68) +where ∆FP[g] is the Jacobian of the transformation +∆FP[g] = det ∂(P1ξ, ˜Λ) +∂(ξ, φ) += det +�P1 +0 +⋆ +1 +� += det P1. +(2.69) +But, one needs to be more careful: +1. The variations involving P1ξ and δti are not orthogonal and, as a consequence, the +measure does not factorize. +2. P1 has zero-modes, i.e. vectors such that P1ξ = 0, which causes the determinant to +vanish, det P1 = 0. +A rigorous analysis will be performed in Section 2.3.2 and will lead to additional factors in +the path integral. +Next, if the actions and measures were invariant under diffeomorphisms and Weyl trans- +formations (which amounts to replace g by ˆg everywhere), it would be possible to factor out +the integrations over the gauge parameters and to cancel the corresponding infinite factors +thanks to the normalization Ωgauge[g]. A new problem arises because the measures are not +Weyl invariant as explained above and one should be careful when replacing the metric +(Section 2.3.3). +4For comparison, Polchinski [193] defines P1 with an overall factor 1/2. +41 + +2.3.2 +Reparametrizations and analysis of P1 +The properties of the operator P1 are responsible for both problems preventing a direct +factorization of the measure; for this reason, it is useful to study it in more details. +The operator P1 is an object which takes a vector v to a symmetric traceless 2-tensor T, +see (2.65a). Conversely, its adjoint P † +1 can be defined from the scalar product (2.34c) +(T, P1v)g = (P † +1 T, v)g, +(2.70) +and takes symmetric traceless tensors to vectors. In components, one finds +(P † +1 T)a = −2∇bTab. +(2.71) +The Riemann–Roch theorem relates the dimension of the kernels of both operators [172]: +dim ker P † +1 − dim ker P1 = −3χg = 6g − 6. +(2.72) +Teichmüller deformations +We first need to characterize Teichmüller deformations, the variations of moduli parameters +which lead to transformations of the metric independent from diffeomorphisms and Weyl +rescalings. This means that the different variations must be orthogonal for the inner-product +(2.34). +First, the deformations must be traceless, otherwise they can be compensated by a Weyl +transformation. The traceless metric variations δg which cannot be generated by a vector +field ξ are perpendicular to P1ξ (otherwise, the former would a linear combination of the +latter): +(δg, P1ξ)g = 0 +=⇒ +(P † +1 δg, ξ)g = 0. +(2.73) +Since ξ is arbitrary, this means that the first argument vanishes +P † +1 δg = 0. +(2.74) +Metric variations induced by a change in the moduli ti are in the kernel of P † +1 +δg ∈ ker P † +1 . +(2.75) +Elements of ker P † +1 are called quadratic differentials and a basis (not necessarily orthonor- +mal) of ker P † +1 is denoted as: +ker P † +1 = Span{φi}, +i = 1, . . . , dim ker P † +1 +(2.76) +(these should not be confused with the Liouville field). The dimension of ker P † +1 is in fact +equal to the dimension of the moduli space (2.51): +dimR ker P † +1 = Mg = +� +� +� +� +� +0 +g = 0, +2 +g = 1, +6g − 6 +g > 1. +(2.77) +The last two terms in the variation (2.64) of δgab are not orthogonal. Let’s introduce +the projector on the complement space of ker P † +1 +Π := P1 +1 +P † +1 P1 +P † +1 . +(2.78) +42 + +The moduli variations can then be rewritten as +δti µi = δti (1 − Π)µi + δti Πµi = δti (1 − Π)µi + δti P1ζi. +(2.79) +The ζi exist because Πµi ∈ Im P1, and they read +ζi := +1 +P † +1 P1 +P † +1 µi. +(2.80) +The first term can be decomposed on the quadratic differential basis (2.76) +(1 − Π)µi = φj(M −1)jk(φk, µi)g +(2.81) +where +Mij := (φi, φj)g. +(2.82) +Ultimately, the variation (2.64) becomes +δgab = (P1 ˜ξ)ab + 2˜Λ gab + Qiab δti. +(2.83) +where +˜ξ = ξ + ζiδti, +Qiab = φjab (M −1)jk(φk, µi)g. +(2.84) +Correspondingly, the norm of the variation splits in three terms since each variation is +orthogonal to the others: +|δg|2 +g = |δ˜Λ| +2 +g + |P1 ˜ξ| +2 +g + |Qiδti|2 +g. +(2.85) +Since the norm is decomposed as a sum, the measure factorizes: +dggab = dg ˜Λ dg(P1 ˜ξ) dg(Qiδti). +(2.86) +One can then perform a change of coordinates +(˜ξ, ˜Λ, Qiδti) −→ (ξ, Λ, δti), +(2.87) +where Λ was defined in (2.65c). The goal of this transformation is to remove the dependence +in the moduli from the measures on the Weyl factor and vector fields, and to recover a finite- +dimensional integral over the moduli: +dg ˜Λ dg(P1 ˜ξ) dg(Qiδti) = dMgt dgΛ dg(P1ξ) det(φi, µj)g +� +det(φi, φj)g +, +(2.88) +where the determinants correspond to the Jacobian. The role of the determinant in the +denominator is to ensure a correct normalization when the basis is not orthonormal (in +particular, it ensures that the Jacobian is independent of the basis). Plugging this result in +(2.28) gives the partition function as +Zg = +� +Tg +dMgt +1 +Ωgauge[ˆg] +� +dgΛ dg(P1ξ) det(φi, µj)g +� +det(φi, φj)g +Zm[g]. +(2.89) +The ti are integrated over the Teichmüller space Tg defined by (2.49) because the vectors ξ +generate only reparametrizations connected to the identity, and thus the remaining freedom +lies in Met(Σg)/G0. Next, we study how to perform the changes of variables to remove P1 +from the measure. +43 + +Conformal Killing vectors +In this section, we focus on the dgΛ dg(P1ξ) part of the measure and we make contact with +the rest at the end. +Infinitesimal reparametrizations generated by a vector field ξa produce only transform- +ations close to the identity. For this reason, integrating over all possible vector fields yields +the volume (2.60a) of the component of the diffeomorphism group connected to the identity: +� +dgξ = ΩDiff0[ˆg]. +(2.90) +Remember that the volume depends only on the moduli, but obviously not on ξ (integrated +over) nor φ (the inner-product (2.34b) is invariant). But, due to the existence of zero-modes, +one gets an integration over a subset of all vector fields, and this complicates the program, +as we discuss now. +Zero-modes ξ(0) of P1 are called conformal Killing vectors (CKV) +ξ(0) ∈ Kg := ker P1 +(2.91) +and satisfy the conformal Killing equation (see also Section 5.1): +(P1ξ(0))ab = ∇aξ(0) +b ++ ∇bξ(0) +a +− gab∇cξ(0)c = 0. +(2.92) +CKVs correspond to reparametrizations which can be absorbed by a change of the con- +formal factor. They should be removed from the ξ integration in order to not double-count +the corresponding metrics. The dimension of the zero-modes CKV space depends on the +genus [172]: +Kg := dimR Kg = dimR ker P1 = +� +� +� +� +� +6 +g = 0, +2 +g = 1, +0 +g > 1. +(2.93) +The associated transformations will be interpreted later (Chapter 5). The groups generated +by the CKVs are +g = 0 : +K0 = SL(2, C), +g = 1 : +K1 = U(1) × U(1). +(2.94) +Note that the first group is non-compact while the second is compact. +A general vector ξ can be separated into a zero-mode part and its orthogonal complement +ξ′: +ξ = ξ(0) + ξ′, +(2.95) +such that +(ξ(0), ξ′)g = 0 +(2.96) +for the inner-product (2.34b). Because zero-modes are annihilated by P1, the correct change +of variables in the partition function (2.66) maps to ξ′ only: +(P1ξ, Λ) −→ (ξ′, φ). +(2.97) +Integrating over ξ at this stage would double count the CKV (since they are already described +by the φ integration). The appropriate Jacobian reads +dgΛ dg(P1ξ) = dgφ dgξ′ ∆FP[g], +(2.98) +where the Faddeev–Popov determinant is +∆FP[g] = det′ ∂(P1ξ, Λ) +∂(ξ′, φ) += det′ P1 = +� +det′ P1P † +1 , +(2.99) +44 + +the prime on the determinant indicating that the zero-modes are excluded. This brings the +partition function (2.89) to the form +Zg = +� +Tg +dMgt Ωgauge[ˆg]−1 +� +dgφ dgξ′ +det(φi, µj)g +� +det(φi, φj)g +∆FP[g]Zm[g]. +(2.100) +Computation – Equation (2.98) +The Jacobian can be evaluated directly: +∆FP[g] = det′ ∂(P1ξ, Λ) +∂(ξ′, φ) += det′ +� P1 +0 +1 +2∇ +1 +� += det′ P1. +(2.101) +As a consequence of det′ P † +1 = det′ P1, the Jacobian can be rewritten as: +� +det′ P † +1 P1 = det′ P1. +(2.102) +It is instructive to derive this result also by manipulating the path integral. Con- +sidering small variations of the fields, one has: +1 = +� +dgδΛ dg(P1δξ) e−|δΛ|2 +g−|P1δξ′|2 +g += ∆FP[g] +� +dgδφ dgδξ′ e−|δφ+ 1 +2 ∇cδξc|2 +g−|P1δξ′|2 +g += ∆FP[g] +� +dgδφ dgδξ′ e−|δφ|2 +g−(δξ′,P † +1 P1δξ′)g += ∆FP[g] +� +det′ P † +1 P1 +�−1/2 +. +That the expression is equal to 1 follows from the normalization of symmetric tensors +and scalars (2.34) (the measures appearing in the path integral (2.89) arises without any +factor). The third equality holds because the measure is invariant under translations +of the fields, and we used the definition of the adjoint. +The volume of the group generated by the vectors orthogonal to the CKV is denoted as +Ω′ +Diff0[g] := Ω′ +Diff0[ˆg] = +� +dgξ′. +(2.103) +As explained in the beginning of this section, one should extract the volume of the full Diff0 +group, not only the volume Ω′ +Diff0[g]. Since the two sets of vectors are orthogonal, we can +expect the measures, and thus the volumes, to factorize. However, a Jacobian can and does +arise: its role it to take into account the normalization of the zero-modes. Denoting by ψi +a basis (not necessarily orthonormal) for the zero-modes +ker P1 = Span{ψi}, +i = 1, . . . , Kg, +(2.104) +the change of variables +ξ′ −→ ξ +(2.105) +reads +dgξ′ = +1 +� +det(ψi, ψj)g +dgξ +Ωckv[g], +(2.106) +45 + +where Ωckv[g] is the volume of the CKV group. The determinant is necessary when the basis +is not orthonormal. The relation between the gauge volumes is then +ΩDiff0[g] = +� +det(ψi, ψj)g Ωckv[g] Ω′ +Diff0[g]. +(2.107) +Note that the CKV volume is given in (2.111) and depends only on the topology but not on +the metric. By using arguments similar to the ones which lead to (2.61), one can expect that +each term is independently invariant under Weyl rescaling: this is indeed true (Section 2.3.3). +Computation – Equation (2.106) +Let’s expand ξ(0) on the zero-mode basis +ξ(0) = αiψi, +(2.108) +where the αi are real numbers, such that one can write the changes of variables +ξ −→ (ξ′, αi). +(2.109) +The Jacobian is computed from +1 = +� +dξ e−|ξ|2 +g = J +� +dξ(0) dξ′ e−|ξ′|2 +g−|ξ(0)| +2 +g += J +� � +i +dαi e−αiαj(ψi,ψj)g +� +dξ′ e−|ξ′|2 +g += J (det(ψi, ψj)g)−1/2 . +Note that the integration over the αi is a standard finite-dimensional integral. This +gives +dξ = +� +det(ψi, ψj)g dξ′ � +i +dαi. +(2.110) +Since nothing depends on the αi, they can be integrated over as in (2.53), giving the +volume of the CKV group +Ωckv[g] = +� � +i +dαi. +(2.111) +Replacing the integration over ξ′ thanks to (2.106), the path integral becomes +Zg = +� +Tg +dMgt Ωgauge[ˆg]−1 +� +dgφ dgξ +det(φi, µj)g +� +det(φi, φj)g +Ωckv[g]−1 +� +det(ψi, ψj)g +∆FP[g] Zm[g]. +(2.112) +Since the matter action and measure, and the Liouville measure are invariant under +reparametrizations, one can perform a change of variables +(f ∗ˆg, f ∗φ, f ∗Ψ) −→ (ˆg, φ, Ψ) +(2.113) +such that everything becomes independent of f (or equivalently ξ). Since the measure for ξ +is Weyl invariant, it is possible to separate it from the rest of the expression, which yields +an overall factor of ΩDiff0[g]. This brings the partition function to the form +Zg = +� +Tg +dMgt ΩDiff0[ˆg] +Ωgauge[ˆg] +� +dgφ det(φi, µj)g +� +det(φi, φj)g +Ωckv[g]−1 +� +det(ψi, ψj)g +∆FP[g] Zm[g] +(2.114) +46 + +where the same symbol is used for the metric +gab := g(φ) +ab = e2φˆgab. +(2.115) +Since the expression is invariant under the full diffeomorphism group Diff(Σg) and not +just under its component Diff0(Σg), one needs to extract the volume of the full diffeomorph- +ism group before cancelling it with the normalization factor. Otherwise, there is still an +over-counting the configurations. Using the relation (2.60c) leads to: +Zg = +1 +ΩΓg +� +Tg +dMgt ΩDiff[ˆg] +Ωgauge[ˆg] +� +dgφ det(φi, µj)g +� +det(φi, φj)g +Ωckv[g]−1 +� +det(ψi, ψj)g +∆FP[g] Zm[g]. +(2.116) +The volume ΩΓg can be factorized outside the integral because it depends only on the genus +and not on the metric. Finally, using the relation (2.52), one can replace the integration +over the Teichmüller space by an integration over the moduli space +Zg = +� +Mg +dMgt ΩDiff[ˆg] +Ωgauge[ˆg] +� +dgφ det(φi, µj)g +� +det(φi, φj)g +Ωckv[g]−1 +� +det(ψi, ψj)g +∆FP[g] Zm[g]. +(2.117) +2.3.3 +Weyl transformations and quantum anomalies +The next question is whether the integrand depends on the Liouville mode φ such that +the Weyl volume can be factorized out. While the matter action has been chosen to be +Weyl invariant – see the condition (2.19) – the measures cannot be defined to be Weyl +invariant. This means that there is a Weyl (or conformal) anomaly, i.e. a violation of the +Weyl invariance due to quantum effects. Since the techniques needed to derive the results +of this section are outside the scope of this book, we simply state the results. +It is possible to show that the Weyl anomaly reads [53, p. 929]5 +∆FP[e2φˆg] +� +det(φi, φj)e2φˆg += e +cgh +6 SL[ˆg,φ] +∆FP[ˆg] +� +det(ˆφi, ˆφj)ˆg +(2.118a) +Zm[e2φˆg] = e +cm +6 SL[ˆg,φ]Zm[ˆg], +(2.118b) +where SL is the Liouville action +SL[ˆg, φ] := 1 +4π +� +d2σ +� +ˆg +� +ˆgab∂aφ∂bφ + ˆRφ +� +, +(2.119) +where ˆR is the Ricci scalar of the metric ˆgab. These relations require to introduce counter- +terms, discussed further in Section 2.3.4. The coefficients cm and cgh are the central charges +respectively of the matter and ghost systems, with: +cgh = −26. +(2.120) +This value will be derived in Section 7.2. +The inner-products between φi and µj, and between the ψi, and the CKV volume are +independent of φ [172, sec. 14.2.2, 53, p. 931] +det(φi, µj)e2φˆg = det(ˆφi, ˆµj)ˆg, +det(ψi, ψj)e2φˆg = det(ψi, ψj)ˆg, +Ωckv[e2φˆg] = Ωckv[ˆg]. +(2.121) +5The relation is written for Zm since the action is invariant and is not affected by the anomaly. +47 + +Remark 2.13 (Weyl and gravitational anomalies) The Weyl anomaly translates into +a non-zero trace of the quantum energy–momentum tensor +⟨gµνTµν⟩ = c +12 R, +(2.122) +where c is the central charge of the theory. The Weyl anomaly can be traded for a gravita- +tional anomaly, which means that diffeomorphisms are broken at the quantum level [122]. +Inserting (2.118) in (2.117) yields +Zg = +� +Mg +dMgt ΩDiff[ˆg] +Ωgauge[ˆg] +det(φi, ˆµj)ˆg +� +det(φi, φj)ˆg +Ωckv[ˆg]−1 +� +det(ψi, ψj)ˆg +∆FP[ˆg] Zm[ˆg] +� +dgφ e− +cL +6 SL[ˆg,φ], +(2.123) +with the Liouville central charge +cL := 26 − cm. +(2.124) +The critical “dimension” is defined to be the value of the matter central charge cm such that +the Liouville central charge cancels +cL = 0 +=⇒ +cm = 26. +(2.125) +If the number of non-compact dimensions is D, it means that the central charge (2.21) of +the transverse CFT satisfies +c⊥ = 26 − D. +(2.126) +In this case, the integrand does not depend on the Liouville mode (because ΩDiff is +invariant under Weyl transformations) and the integration over φ can be factored out and +yields the volume of the Weyl group (2.60b) +� +dgφ = ΩWeyl[ˆg]. +(2.127) +Then, taking +Ωgauge[ˆg] = ΩDiff[ˆg] × ΩWeyl[ˆg] +(2.128) +removes the infinite gauge contributions and gives the partition function +Zg = +� +Mg +dMgt det(φi, ˆµj)ˆg +� +det(φi, φj)ˆg +Ωckv[ˆg]−1 +� +det(ψi, ψj)ˆg +∆FP[ˆg] Zm[ˆg]. +(2.129) +2.3.4 +Ambiguities, ultralocality and cosmological constant +Different ambiguities remain in the previous computations, starting with the definitions of +the measures (2.32) and (2.34), then in obtaining the volume of the diffeomorphism (2.60a) +and Weyl (2.60b) groups, and finally in deriving the conformal anomaly (2.118). +These different ambiguities can be removed by renormalizing the worldsheet cosmological +constant. This implies that the action +Sµ[g] = +� +d2σ√g +(2.130) +must be added to the classical Lagrangian, where µ0 is the bare cosmological constant. This +means that Weyl invariance is explicitly broken at the classical level. After performing all +the manipulations, µ0 is determined by removing all ambiguities and enforcing invariance +under the Weyl symmetry at the quantum level. +This amounts to set the renormalized +cosmological constant to zero (since it breaks the Weyl symmetry). +The possibility to +48 + +introduce a counter-term violating a classical symmetry arises because the symmetry itself +is broken by a quantum anomaly, so there is no reason to enforce it in the classical action. +We now review each issue separately. First, consider the inner-product of a single tensor +(2.32): the determinant det γg depends on the metric and one should be more careful when +fixing the gauge or integrating over all metrics. +However, ultralocality implies that the +determinant can only be of the form [53, pp. 923] +� +det γg = e−µγ Sµ[g], +(2.131) +for some µγ ∈ R, since Sµ is the only renormalizable covariant functional depending on the +metric but not on its derivatives. The effect is just to redefine the cosmological constant. +Second, the volume of the field space can be defined as the limit λ → 0 of a Gaussian +integral [53, pp. 931]: +ΩΦ = lim +λ→0 +� +dgΦ e−λ (Φ,Φ)g. +(2.132) +Due to ultralocality, the Gaussian integral should again be of the form +� +dgΦ e−λ (Φ,Φ)g = e−µ(λ) Sµ[g], +(2.133) +for some constant µ(λ). Hence, the limit λ → 0 gives +ΩΦ = +� +dgΦ = e−µ(0) Sµ[g], +(2.134) +which can be absorbed in the cosmological constant. However, the situation is more com- +plicated if Φ = ξ, φ since the integration variables also appear in the measure, as it was +also discussed before (2.61). But, in that case, it cannot appear in the expression of the +volume in the LHS. Moreover, invariances under diffeomorphisms for both measures, and +under Weyl rescalings for the vector measure, imply that the LHS can only depend on the +moduli through the background metric ˆg. The diffeomorphism and Weyl volumes can be +written in terms of e−ˆµ Sµ[ˆg]: since there is no counter-term left (the cosmological constant +counter-term is already fixed to cancel the coefficient of Sµ[g]), it is necessary to divide by +Ωgauge to cancel the volumes. +Finally, the computation of the Weyl anomaly (2.118) yields divergent terms of the form +lim +ϵ→0 +1 +ϵ +� +d2σ√g. +(2.135) +These divergences are canceled by the cosmological constant counter-term, see [54, app. 5.A] +for more details. +2.3.5 +Gauge-fixed path integral +As a conclusion of this section, we found that the partition function (2.28) can be written +as +Zg = +� +Mg +dMgt det(φi, ˆµj)ˆg +� +det(φi, φj)ˆg +Ωckv[ˆg]−1 +� +det(ψi, ψj)ˆg +∆FP[ˆg] Zm[ˆg], +(2.136a) += +� +Mg +dMgt +� +det(φi, ˆµj)2 +ˆg +det(φi, φj)ˆg +det′ ˆP † +1 ˆP1 +det(ψi, ψj)ˆg +Zm[ˆg] +Ωckv[ˆg]. +(2.136b) +after gauge fixing of the worldsheet diffeomorphisms and Weyl rescalings. It is implicit that +the factors for the CKV and moduli are respectively absent for g > 1 and g < 1. For g = 0 +the CKV group is non-compact and its volume is infinite. It looks like the partition vanishes, +but there are subtleties which will be discussed in Section 3.1.3. +49 + +Remark 2.14 (Weil–Petersson metric) When the metric is chosen to be of constant +curvature ˆR = −1, the moduli measure together with the determinants form the Weil– +Petersson measure +d(WP) = +� +Mg +dMgt det(φi, ˆµj)ˆg +� +det(φi, φj)ˆg +. +(2.137) +In (2.136), the background metric ˆgab is fixed. However, the derivation holds for any +choice of ˆgab: as a consequence, it makes sense to relax the gauge fixing and allow it to vary +while adding gauge symmetries. The first symmetry is background diffeomorphisms: +σ′a = ˆf a(σb), +ˆg′(σ′) = f ∗ˆg(σ), +φ′(σ′) = f ∗φ(σ), +Ψ′(σ′) = f ∗Ψ(σ). +(2.138) +This symmetry is automatic for Sm[ˆg, Ψ] since Sm[g, Ψ] was invariant under (2.5). Similarly, +the integration measures are also invariant. A second symmetry is found by inspecting the +decomposition (2.56) +gab = f ∗� +e2φˆgab(t) +� +, +(2.139) +which is left invariant under a background Weyl symmetry (also called emergent): +g′ +ab(σ) = e2ω(σ)gab(σ), +φ′(σ) = φ(σ) − ω(σ), +Ψ′(σ) = Ψ(σ). +(2.140) +Let us stress that it is not related to the Weyl rescaling (2.10) of the metric gab. +The +background Weyl rescaling (2.140) is a symmetry even when the physical Weyl rescaling +(2.10) is not. Together, the background diffeomorphisms and Weyl symmetry have three +gauge parameters, which is sufficient to completely fix the background metric ˆg up to moduli. +In fact, the combination of both symmetries is equivalent to invariance under the physical +diffeomorphisms. +To prove this statement, consider two metrics g and g′ related by a +diffeomorphism F and both gauge fixed to pairs (f, φ, ˆg) and (f ′, φ′, ˆg′): +g′ +ab = F ∗gab, +g′ +ab = f ′∗� +e2φ′ˆg′ +ab +� +, +gab = f ∗� +e2φˆgab +� +. +(2.141) +Then, the gauge fixing parametrizations are related by background symmetries ( ˆF, ω) as +ˆF = f ′−1 ◦ F ◦ f, +φ′ = ˆF ∗(φ − ω), +ˆg′ +ab = ˆF ∗(e2ωˆgab). +(2.142) +Moreover, this also implies that there is a diffeomorphism ˜f = F ◦ f such that g′ is gauge +fixed in terms of (φ, ˆg): +g′ +ab = ˜f ∗� +e2φˆgab +� +. +(2.143) +Computation – Equation (2.142) +The functions F, f, f ′, φ, φ′ and the metrics gab, g′ +ab, ˆgab and ˆg′ +ab are all fixed and one +must find ˆF and ω such that the relations (2.141) are compatible. First, one rewrites +g′ +ab in terms of ˆgab and compare with the expression with ˆg′ +ab: +g′ +ab = F ∗gab = F ∗� +f ∗� +e2φˆgab +�� += F ∗� +f ∗� +e2(φ−ω)e2ωˆgab +�� += f ′∗� +e2φ′ˆg′ +ab +� +. +In the third equality, we have introduced ω because ˆg′ +ab = ˆF ∗ˆgab is not true in general +since there are 3 independent components but ˆF has only 2 parameters, so we cannot +just define f ′ = F ◦ f and φ′ = φ. This explains the importance of the emergent Weyl +symmetry. +50 + +Remark 2.15 (Gauge fixing and field redefinition) Although it looks like we are un- +doing the gauge fixing, this is not exactly the case since the original metric is not used any- +more. One can understand the procedure of this section as a field redefinition: the degrees of +freedom in gab are repackaged into two fields (φ, ˆgab) adapted to make some properties of the +system more salient. A new gauge symmetry is introduced to maintain the number of degrees +of freedom. The latter helps to understand the structure of the action on the background. +Finally, in this context, the Liouville action is understood as a Wess–Zumino action, which +is defined as the difference between the effective actions evaluated in each metric. Another +typical use of this point of view is to rewrite a massive vector field as a massless gauge field +together with an axion [195]. +Remark 2.16 (Two-dimensional gravity) In 2d gravity, one does not work in the crit- +ical dimension (2.125) and cL ̸= 0. Thus, the Liouville mode does not decouple: the con- +formal anomaly breaks the Weyl symmetry at the quantum level which gives dynamics to +gravity, even if it has no degree of freedom classically. +As a consequence, one chooses +Ωgauge = ΩDiff. +Since the role of the classical Weyl symmetry is not as important as for string theory, it is +even not necessary to impose it classically. This leads to consider non-conformal matter [21, +22, 31, 82, 83]. Following the arguments from Section 2.1, the existence of the emergent +Weyl symmetry (2.140) implies that the total action Sgrav[ˆg, φ] + Sm[ˆg, Ψ] must be a CFT +for a flat background ˆg = δ, even if the two actions are not independently CFTs. +2.4 +Ghost action +2.4.1 +Actions and equations of motion +It is well-known that a determinant can be represented with two anticommuting fields, called +ghosts. The fields carry indices dictated by the map induced by the operator of the Faddeev– +Popov determinant: one needs a symmetric and traceless anti-ghost bab and a vector ghost +ca fields: +∆FP[g] = +� +d′ +gb d′ +gc e−Sgh[g,b,c], +(2.144) +where the prime indicates that the ghost zero-modes are omitted. The ghost action is +Sgh[g, b, c] := 1 +4π +� +d2σ√g gabgcdbac(P1c)bd +(2.145a) += 1 +4π +� +d2σ√g gab� +bac∇bcc + bbc∇acc − bab∇ccc� +. +(2.145b) +The ghosts ca and anti-ghosts bab are associated respectively to the variations due to the +diffeomorphisms ξa and to the variations perpendicular to the gauge slice. The normalization +of 1/4π is conventional. In Minkowski signature, the action is multiplied by a factor i. +Since bab is traceless, the last term of the action vanishes and could be removed. However, +this implies to consider traceless variations of the bab when varying the action (to compute +the equations of motion, the energy–momentum tensor, etc.). On the other hand, one can +keep the term and consider unconstrained variation of bab (since the structure of the action +will force the variation to have the correct symmetry), which is simpler. A last possibility is +to introduce a Lagrange multiplier. These aspects are related to the question of introducing +a ghost for the Weyl symmetry, which is described in Section 2.4.2. +The equations of motion are +(P1c)ab = ∇acb + ∇bca − gab∇ccc = 0, +(P † +1 b)a = −2∇bbab = 0. +(2.146) +51 + +Hence, the classical solutions of b and c are respectively mapped to the zero-modes of the +operators P † +1 and P1, and they are thus associated to the CKV and Teichmüller parameters. +The energy–momentum tensor is +T gh +ab = −bac∇bcc − bbc∇acc + cc∇cbab + gabbcd∇ccd. +(2.147) +Its trace vanishes off-shell (i.e. without using the b and c equations of motion) +gabT gh +ab = 0, +(2.148) +which shows that the action (2.145) is invariant under Weyl transformations +Sgh[e2ωg, b, c] = Sgh[g, b, c]. +(2.149) +The action (2.145) also has a U(1) global symmetry. The associated conserved charge is +called the ghost number and counts the number of c ghosts minus the number of b ghosts, +i.e. +Ngh(b) = −1, +Ngh(c) = 1. +(2.150a) +The matter fields are inert under this symmetry: +Ngh(Ψ) = 0. +(2.150b) +In terms of actions, the path integral (2.136) can be rewritten as +Zg = +� +Mg +dMgt det(φi, ˆµj)ˆg +� +det(φi, φj)ˆg +Ωckv[ˆg]−1 +� +det(ψi, ψj)ˆg +� +dˆgΨ d′ +ˆgb d′ +ˆgc e−Sm[ˆg,Ψ]−Sgh[ˆg,b,c]. +(2.151) +One can use (2.136) or (2.151) indifferently: the first is more appropriate when using spectral +analysis to compute the determinant explicitly, while the second is more natural in the +context of CFTs. +2.4.2 +Weyl ghost +Ghosts have been introduced for the reparametrizations (generated by ξa) and the traceless +part of the metric (the gauge field associated to the transformation): one may wonder why +there is not a ghost cw associated to the Weyl symmetry along with an antighost for the trace +of the metric (i.e. the conformal factor). This can be understood from several viewpoints. +First, the relation between a metric and its transformation – and the corresponding gauge +fixing condition – does not involve any derivative: as such, the Jacobian is trivial. Second, +one could choose +F ⊥ +ab = √ggab − +� +ˆgˆgab = 0 +(2.152) +as a gauge fixing condition instead of (2.63), and the trace component does not appear +anywhere. Finally, a local Weyl symmetry is not independent from the diffeomorphisms. +Remark 2.17 (Local Weyl symmetry) The topic of obtaining a local Weyl symmetry by +gauging a global Weyl symmetry (dilatation) is very interesting [86, chap. 15, 113]. Under +general conditions, one can express the new action in terms of the Ricci tensor (or of the +curvature): this means that the Weyl gauge field and its curvature are composite fields. +Moreover, one finds that local Weyl invariance leads to an off-shell condition while diffeo- +morphisms give on-shell conditions. This explains why one imposes only Virasoro constraints +(associated to reparametrizations) and no constraints for the Weyl symmetry in the covariant +quantization. +52 + +However, it can be useful to introduce a ghost field cw for the Weyl symmetry nonetheless. +In view of the previous discussion, this field should appear as a Lagrange multiplier which +ensures that bab is traceless. Starting from the action (2.145), one finds +S′ +gh[g, b, c, cw] = 1 +4π +� +d2σ√g gab� +bac∇bcc + bbc∇acc + 2babcw +� +, +(2.153) +where bab is not traceless anymore. The ghost cw is not dynamical since the action does not +contain derivatives of it, and it can be integrated out of the path integral to recover (2.145). +The equations of motion for this modified action are +∇acb + ∇bca + 2gabcw = 0, +∇abab = 0, +gabbab = 0. +(2.154) +Contracting the first equation with the metric gives +cw = −1 +2 ∇aca, +(2.155) +and thus cw is nothing else than the divergence of the ca field: the Weyl ghost is a composite +field (this makes connection with Remark 2.17) – see also (2.65c). The energy–momentum +tensor of the ghosts with action (2.153) is +T ′gh +ab = − +� +bac∇bcc + bbc∇acc + 2babcw +� +− ∇c(babcc) ++ 1 +2 gabgcd� +bce∇dce + bde∇cce + 2bcdcw +� +. +(2.156) +The trace of this tensor +gabT ′gh +ab = −gab∇c(babcc) +(2.157) +does not vanish off-shell, but it does on-shell since gabbab = 0. This implies that the theory +is Weyl invariant even if the action is not. It is interesting to contrast this with the trace +(2.148) when the Weyl ghost has been integrated out. +The equations of motion (2.146) and energy–momentum tensor (2.147) for the action +(2.145) can be easily derived by replacing cw by its solution in the previous formulas. +Computation – Equation (2.156) +The first parenthesis comes from varying gab, the second from the covariant derivatives, +the last from the √g. The second term comes from +gab� +bacδ∇bcc + bbcδ∇acc� += 2gabbacδ∇bcc = 2gabbacδΓc +bdcd += gabbacgce� +∇bδgde + ∇dδgbe − ∇eδgbd +� +cd += bab� +∇aδgbc + ∇cδgab − ∇bδgac +� +cc += bab∇cδgabcc, +where two terms have cancelled due to the symmetry of bab. Integrating by part gives +the term in the previous equation. +Note that the integration on the Weyl ghost yields a delta function +� +dgcw e−(cw,gabbab)g = δ +� +gabbab +� +. +(2.158) +53 + +2.4.3 +Zero-modes +The path integral (2.151) excludes the zero-modes of the ghosts. One can expect them to +be related to the determinants of elements of ker P1 and ker P † +1 with Grassmann coefficients. +They can be included after few simple manipulations (see also Appendix C.1.3). +It is simpler to first focus on the b ghost (to avoid the problems related to the CKV). +The path integral (2.151) can be rewritten as +Zg = +� +Mg +dMgt +Ωckv[ˆg]−1 +� +det(ψi, ψj)ˆg +� +dˆgΨ dˆgb d′ +ˆgc +Mg +� +i=1 +(b, ˆµi)ˆg e−Sm[ˆg,Ψ]−Sgh[ˆg,b,c]. +(2.159) +In this expression, c zero-modes are not integrated over, only the b zero-modes are. This is +the standard starting point on Riemann surfaces with genus g ≥ 1. The inner-product reads +explicitly +(b, ˆµi)ˆg = +� +d2σ +� +ˆg Gabcd +⊥ +babˆµi,cd = +� +d2σ +� +ˆg gacgbdbabˆµi,cd. +(2.160) +Computation – Equation (2.159) +Since the zero-modes of b are in the kernel of P † +1 , it means that the quadratic differentials +(2.76) also provide a suitable basis: +b = b0 + b′, +b0 = b0iφi, +where the b0i are Grassmann-odd coefficients. The first step is to find the Jacobian for +the changes of variables b → (b′, b0i): +1 = +� +dˆgb e−|b|2 +ˆg = J +� +dˆgb′ � +i +db0i e−|b′|2 +ˆg−|b0iφi|2 = J +� +det(φi, φj). +Next, (2.151) has no zero-modes, so one must insert Mg of them at arbitrary positions +σ0 +j to get a non-vanishing result when integrating over dMgb0i. The result of the integral +is: +� +dMgb0i +� +j +b0(σ0 +j ) = +� +dMgb0i +� +j +� +b0iφi(σ0 +j ) +� += det φi(σ0 +j ). +The only combination of the φi which does not vanish is the determinant due to the +anti-symmetry of the Grassmann numbers. Combining both results leads to: +dˆgb′ +� +det(φi, φj)ˆg += +dˆgb +det φi(σ0 +j ) +Mg +� +j=1 +b(σ0 +j ). +(2.161) +The locations positions σ0 +j are arbitrary (in particular, the RHS does not depend on +them since the LHS does not either). +Note that more details are provided in Ap- +pendix C.1.3. +An even simpler result can be obtained by combining the previous formula with the +factor det(φi, ˆµj)ˆg: +dˆgb′ +det(φi, ˆµj)ˆg +� +det(φi, φj)ˆg += dˆgb +Mg +� +j=1 +(b, ˆµj)ˆg. +(2.162) +54 + +This follows from +Mg +� +j=1 +b(σ0 +j ) = +Mg +� +j=1 +� +b0iφi(σ0 +j ) +� += det φi(σ0 +j ) +Mg +� +j=1 +b0i, +det(φi, ˆµj)ˆg +Mg +� +j=1 +b0i = +Mg +� +j=1 +� +b0i(φi, ˆµj)ˆg +� += +Mg +� +j=1 +(b0iφi, ˆµj)ˆg = +Mg +� +j=1 +(b, ˆµj)ˆg. +Note that the previous manipulations are slightly formal: the symmetric traceless fields +bab and φi,ab carry indices and there should be a product over the (two) independent +components. This is a trivial extension and would just make the notations heavier. +Similar manipulations lead to a new expression which includes also the c zero-mode (but +which is not very illuminating): +Zg = +� +Mg +dMgt Ωckv[ˆg]−1 +det ψi(σ0 +j ) +� +dˆgΨ dˆgb dˆgc +Kc +g +� +j=1 +ϵab +2 ca(σ0 +j )cb(σ0 +j ) +× +Mg +� +i=1 +(ˆµi, b)ˆg e−Sm[ˆg,Ψ]−Sgh[ˆg,b,c]. +(2.163) +The σ0a +j +are Kc +g = Kg/2 fixed positions and the integral does not depend on their values. +Note that only Kc +g positions are needed because the coordinate is 2-dimensional: fixing 3 +points with 2 components correctly gives 6 constraints. Then, ψi(σ0a +j ) is a 6-dimensional +matrix, with the rows indexed by i and the columns by the pair (a, j). +The expression cannot be simplified further because the CKV factor is infinite for g = 0. +This is connected to a fact mentioned previously: there is a remaining gauge symmetry +which is not taken into account +c −→ c + c0, +P1c0 = 0. +(2.164) +A proper account requires to gauge fix this symmetry: the simplest possibility is to insert +three or more vertex operators – this topic is discussed in Section 3.1. +Finally, note that the same question arises for the b-ghost since one has the symmetry +b −→ b + b0, +P † +1 b0 = 0. +(2.165) +That there is no problem in this case is related to the presence of the moduli. +2.5 +Normalization +In the previous sections, the closed string coupling constant gs did not appear in the ex- +pressions. Loops in vacuum amplitudes are generated by splitting of closed strings. By +inspecting the amplitudes, it seems that there are 2g such splittings (Figure 2.2), which +would lead to a factor g2g +s . However, this is not quite correct: this result holds for a 2-point +function. Gluing the two external legs to get a partition function (that is, taking the trace) +leads to an additional factor g−2 +s +(to be determined later), such that the overall factor is +g2g−2 +s +. The fact that it is the appropriate power of the coupling constant can be more easily +understood by considering n-point amplitudes (Section 3.1). The normalization of the path +integral can be completely fixed by unitarity [193]. +The above factor has a nice geometrical interpretation. Defining +Φ0 = ln gs +(2.166) +55 + +and remembering the expression (2.4) of the Euler characteristics χg = 2 − 2g, the coupling +factor can be rewritten as +g2g−2 +s += e−Φ0χg = exp +� +−Φ0 +4π +� +d2σ√gR +� += e−Φ0SEH[g], +(2.167) +where SEH is the Einstein–Hilbert action. +This action is topological in two dimensions. +Hence, the coupling constant can be inserted in the path integral simply by shifting the +action by the above term. +This shows that string theory on a flat target spacetime is +completely equivalent to matter minimally coupled to Einstein–Hilbert gravity with a cos- +mological constant (tuned to impose Weyl invariance at the quantum level). The advantage +of describing the coupling power in this fashion is that it directly generalizes to scattering +amplitudes and to open strings. The parameter Φ0 is interpreted as the expectation value of +the dilaton. Replacing it by a general field Φ(Xµ) is a generalization of the matter non-linear +sigma model, but this topic is beyond the scope of this book. +Figure 2.2: g-loop partition function. +2.6 +Summary +In this chapter, we started with a fairly general matter CFT – containing at least D scalar +fields Xµ – and explained under which condition it describes a string theory. The most +important consequence is that the matter 2d QFT must in fact be a 2d CFT. We then +continued by describing how to gauge fix the integration over the surfaces and we identified +the remaining degrees of freedom – the moduli space Mg – up to some residual redundancy +– the conformal Killing vector (CKV). Then, we showed how to rewrite the result in terms +of ghosts and proved that they are also a CFT. This means that a string theory can be com- +pletely described by two decoupled CFTs: a universal ghost CFT and a theory-dependent +matter CFT describing the string spacetime embedding and the internal structure. The +advantage is that one can forget the path integral formalism altogether and employ only +CFT techniques to perform the computations. This point of view will be developed for off- +shell amplitudes (Chapter 11) in order to provide an alternative description of how to build +amplitudes. It is particularly fruitful because one can also consider matter CFTs which do +not have a Lagrangian description. In the next chapter, we describe scattering amplitudes. +2.7 +Suggested readings +Numerous books have been published on the worldsheet string theory. +Useful (but not +required) complements to this chapter and subsequent ones are [151, 265] for introductory +texts and [24, 47, 48, 128, 193] for more advanced aspects. +• The definition of a field measure from a Gaussian integral and manipulations thereof +can be found in [100, sec. 15.1, 22.1, 172, chap. 14, 191, 53]. +56 + +• The most complete explanations of the gauge fixing procedure are [100, sec. 15.1, 22.1, +24, sec. 3.4, 6.2, 193, chap. 5, 48, 124, chap. 5]. The original derivation can be found +in [52, 161]. +• For the geometry of the moduli space, see [172, 173]. +• Ultralocality and its consequences are described in [53, 191] (see also [98, sec. 2.4]). +• The use of a Weyl ghost is shown in [240, sec. 8, 258, sec. 9.2]. +57 + +Chapter 3 +Worldsheet path integral: +scattering amplitudes +Abstract +In this chapter, we generalize the worldsheet path integral to compute scattering +amplitudes, which corresponds to insert vertex operators. The gauge fixing from the previous +chapter is generalized to this case. In particular, we discuss the 2-point amplitude on the +sphere. Finally, we introduce the BRST symmetry and motivate some properties of the +BRST quantization, which will be performed in details later. The formulas in this chapter +are all covariant: they will be rewritten in complex coordinates in the next chapter. +3.1 +Scattering amplitudes on moduli space +In this section, we describe the scattering of n strings. The momentum representation is +more natural for describing interactions, especially in string theory. Therefore, each string is +characterized by a state Vαi(ki) with momentum ki and some additional quantum numbers +αi (i = 1, . . . , n). We start from the worldsheet path integral (2.28) before gauge fixing: +Zg = +� +dggab +Ωgauge[g] Zm[g], +Zm[g] = +� +dgΨ e−Sm[g,Ψ]. +(3.1) +3.1.1 +Vertex operators and path integral +The external states are represented by infinite semi-tubes attached to the surfaces. Under a +conformal mapping, the tubes can be mapped to points called punctures on the worldsheet. +At g loops, the resulting space is a Riemann surface Σg,n of genus g with n punctures (or +marked points). The external states are represented by integrated vertex operators +Vα(ki) := +� +d2σ +� +g(σ) Vα(k; σ). +(3.2) +The vertex operators Vα(k; σ) are built from the matter CFT operators and from the world- +sheet metric gab. The functional dependence is omitted to not overload the notation, but +one should read Vα(k; σ) := Vα[g, Ψ](k; σ). The integration over the state positions is neces- +sary because the mapping of the tube to a point is arbitrary. Another viewpoint is that it +is needed to obtain an expression invariant under worldsheet diffeomorphisms. The vertex +operators described general states which not necessarily on-shell: this restriction will be +found later when discussing the BRST invariance of scattering amplitudes (Section 3.2.2). +58 + +Following Section 2.3.5, the Einstein–Hilbert action with boundary term +SEH[g] := 1 +4π +� +d2σ√g R + 1 +2π +� +ds k = χg,n. +(3.3) +is inserted in the path integral equals the Euler characteristics χg,n (the g in χg,n denotes +the genus). On a surface with punctures, the latter is shifted by the number of punctures +(which are equivalent to boundaries or disks) with respect to (2.4): +χg,n := χ(Σg,n) = 2 − 2g − n. +(3.4) +This gives the normalization factor: +g−χg,n +s += e−Φ0SEH[g], +Φ0 := ln gs. +(3.5) +The correctness factor can be verified by inspection of the Riemann surface for the scat- +tering of n string at g loops. In particular, the string coupling constant is by definition +the interaction strength for the scattering of 3 strings at tree-level. Moreover, the tree-level +2-point amplitude contains no interaction and should have no power of gs. This factor can +also be obtained by unitarity [193]. +By inserting these factors in (2.28), the g-loop n-point scattering amplitude is described +by: +Ag,n({ki}){αi} := +� +dggab +Ωgauge[g] dgΨ e−Sm[g,Ψ]−Φ0SEH[g] +n +� +i=1 +�� +d2σi +� +g(σi) Vαi(ki; σi) +� +. +(3.6) +The σi dependence of each √g will be omitted from now on since no confusion is possible. +The following equivalent notations will be used: +Ag,n({ki}){αi} := Ag,n(k1, . . . , kn)α1,...,αn := Ag,n +� +Vα1(k1), . . . , Vαn(kn) +� +. +(3.7) +The complete (perturbative) amplitude is found by summing over all genus: +An(k1, . . . , kn)α1,...,αn = +∞ +� +g=0 +Ag,n(k1, . . . , kn)α1,...,αn. +(3.8) +We omit a genus-dependent normalization which can be determined from unitarity [193]. +Sometimes, it is convenient to extract the factor e−Φ0χg,n of the amplitude Ag,n to display +explicitly the genus expansion, but we will not follow this convention here. Since each term of +the sum scales as Ag,n ∝ g2g+n−2 +s +, this expression clearly shows that worldsheet amplitudes +are perturbative by definition: this motivates the construction of a string field theory from +which the full non-perturbative S-matrix can theoretically be computed. +Finally, the amplitude (3.6) can be rewritten in terms of correlation functions of the +matter QFT integrated over worldsheet metrics: +Ag,n({ki}){αi} = +� +dggab +Ωgauge[g] e−Φ0SEH[g] +� +n +� +i=1 +d2σi +√g +� n +� +i=1 +Vαi(ki; σi) +� +m,g +. +(3.9) +The correlation function plays the same role as the partition function in (2.28). This shows +that string expressions are integrals of CFT expressions over the space of worldsheet metrics +(to be reduced to the moduli space). +We address a last question before performing the gauge fixing: what does (3.6) computes +exactly: on-shell or off-shell? Green functions or amplitudes? if amplitudes, the S-matrix +59 + +or just the interacting part T (amputated Green functions)? The first point is that a path +integral over connected worldsheets will compute connected processes. We will prove later, +when discussing the BRST quantization, that string states must be on-shell (Sections 3.2 +and 3.2.2) and that it corresponds to setting the Hamiltonian (2.26) to zero: +H = 0. +(3.10) +From this fact, it follows that (2.28) must compute amplitudes since non-amputated Green +functions diverge on-shell (due to external propagators). Finally, the question of whether it +computes the S-matrix S = 1 + iT, or just the interacting part T is subtler. At tree-level, +they agree for n ≥ 3, while T = 0 for n = 2 and S reduces to the identity. This difficulty +(discussed further in Section 3.1.2) is thus related to the question of gauge-fixing tree-level +2-point amplitude (Section 3.1.3). It has long been believed that (2.28) computes only the +interacting part (amputated Green functions), but it has been understood recently that this +is not correct and that (2.28) computes the S-matrix. +Remark 3.1 (Scattering amplitudes in QFT) Remember that the S-matrix is separ- +ated as: +S = 1 + iT, +(3.11) +where 1 denotes the contribution where all particles propagate without interaction. +The +connected components of S and T are denoted by Sc and T c. +The n-point (connected) +scattering amplitudes An for n ≥ 3 can be computed from the Green functions Gn through +the LSZ prescription (amputation of the external propagators): +An(k1, . . . , kn) = Gn(k1, . . . , kn) +n +� +i=1 +(k2 +i + m2 +i ). +(3.12) +The path integral computes the Green functions Gn; perturbatively, they are obtained from +the Feynman rules. They include a D-dimensional delta function +Gn(k1, . . . , kn) ∝ δ(D)(k1 + · · · + kn). +(3.13) +The 2-point amputated Green function T2 computed from the LSZ prescription vanishes +on-shell. For example, considering a scalar field at tree-level, one finds: +T2 = G2(k, k′) (k2 + m2)2 ∼ (k2 + m2) δ(D)(k + k′) −−−−−−→ +k2→−m2 0 +(3.14) +since +G2(k, k′) = δ(D)(k + k′) +k2 + m2 +. +(3.15) +Hence, T2 = 0 and the S-matrix (3.11) reduces to the identity component Sc +2 = 12 (which is +a connected process). There are several way to understand this result: +1. The recursive definition of the connected S-matrix Sc from the cluster decomposition +principle requires a non-vanishing 2-point amplitude [121, sec. 5.1.5, 251, sec. 4.3, 63, +sec. 6.1]. +2. The 2-point amplitude corresponds to the normalization of the 1-particle states (overlap +of a particle state with itself, which is non-trivial) [250, eq. 4.1.4, 239, chap. 5]. +3. A single particle in the far past propagating to the far future without interacting is a +connected and physical process [63, p. 133]. +4. It is required by the unitarity of the 2-point amplitude [66]. +60 + +These points indicate that the 2-point amplitude is proportional to the identity in the mo- +mentum representation [121, p. 212, 250, eq. 4.3.3 and 4.1.5] +A2(k, k′) = 2k0 (2π)D−1δ(D−1)(k − k′). +(3.16) +The absence of interactions implies that the spatial momentum does not change (the on- +shell condition implies that the energy is also conserved). This relation is consistent with +the commutation relation of the operators with the Lorentz invariant measure1 +[a(k), a†(k′)] = 2k0 (2π)D−1δ(D−1)(k − k′). +(3.17) +That this holds for all particles at all loops can be proven using the Källen–Lehman repres- +entation [121, p. 212]. +On the other hand, the identity part in (3.11) is absent for n ≥ 3 for connected amp- +litudes: Sc +n = T c +n for n ≥ 3. This shows that the Feynman rules and the LSZ prescription +compute only the interacting part T of the on-shell scattering amplitudes. The reason is that +the derivation of the LSZ formula assumes that the incoming and outgoing states have no +overlap, which is not the case for the 2-point function. A complete derivation of the S-matrix +from the path integral is more involved [121, sec. 5.1.5, 260, sec. 6.7, 81] (see also [37]). +The main idea is to consider a superposition of momentum states (here, in the holomorphic +representation [260, sec. 5.1, 6.4]) +φ(α) = +� +dD−1k α(k)∗a†(k). +(3.18) +They contribute a quadratic piece to the connected S-matrix and, setting them to delta func- +tions, one recovers the above result. +3.1.2 +Gauge fixing: general case +The Faddeev–Popov gauge fixing of the worldsheet diffeomorphisms and Weyl rescaling +(2.15) goes through also in this case if the integrated vertex operators are diffeomorphism +and Weyl invariant: +δξVαi(ki) = δξ +� +d2σ√g Vαi(ki; σ) = 0, +(3.19a) +δωVαi(ki) = δω +� +d2σ√g Vαi(ki; σ) = 0, +(3.19b) +with the variations defined in (2.7) and (2.11). Diffeomorphism invariance is straightforward +if the states are integrated worldsheet scalars. However, if the states are classically Weyl +invariant, they are not necessary so at the quantum level: vertex operators are composite +operators, which need to be renormalized to be well-defined at the quantum level. Renor- +malization introduces a scale which breaks Weyl invariance. Enforcing it to be a symmetry +of the vertex operators leads to constraints on the latter. We will not enter in the details +since it depends on the matter CFT and we will assume that the operators Vαi(ki) are indeed +Weyl invariant (see [193, sec. 3.6] for more details). In the rest of this book, we will use +CFT techniques developed in Chapter 6. The Einstein–Hilbert action is clearly invariant +under both symmetries since it is a topological quantity. +1If the modes are defined as ˜a(k) = a(k)/ +√ +2k0 such that [˜a(k), ˜a†(k′)] = (2π)D−1δ(D−1)(k − k′), then +one finds ˜ +A2(k, k′) = (2π)D−1δ(D−1)(k − k′). +61 + +Following the computations from Section 2.3 leads to a generalization of (2.136) with +the vertex operators inserted for the amplitude (3.6): +Ag,n({ki}){αi} = g−χg,n +s +� +Mg +dMgt det(φi, ˆµj)ˆg +� +det(φi, φj)ˆg +Ωckv[ˆg]−1 +� +det(ψi, ψj)ˆg +× +� +n +� +i=1 +d2σi +� +ˆg +� n +� +i=1 +ˆVαi(ki; σi) +� +m,ˆg +. +(3.20) +The hat on the vertex operators indicates that they are evaluated in the background metric +ˆg. +The next step is to introduce the ghosts: following Section 2.4, the generalization of +(2.159) is +Ag,n({ki}){αi} = g−χg,n +s +� +Mg +dMgt +Ωckv[ˆg]−1 +� +det(ψi, ψj)ˆg +� +dˆgb d′ +ˆgc +Mg +� +i=1 +(b, ˆµi)ˆg e−Sgh[ˆg,b,c] +× +� +n +� +i=1 +d2σi +� +ˆg +� n +� +i=1 +ˆVαi(ki; σi) +� +m,ˆg +. +(3.21) +For the moment, only the b ghosts come with zero-modes. +Then, c zero-modes can be +introduced in (3.21) +Ag,n = g−χg,n +s +� +Mg +dMgt Ωckv[ˆg]−1 +det ψi(σ0 +j ) +� +dˆgb dˆgc +Kc +g +� +j=1 +ϵab +2 ca(σ0 +j )cb(σ0 +j ) +Mg +� +i=1 +(ˆµi, b)ˆg e−Sgh[ˆg,b,c] +× +� +n +� +i=1 +d2σi +� +ˆg +� n +� +i=1 +ˆVαi(ki; σi) +� +m,ˆg +, +(3.22) +by following the same derivation as (2.163). The formulas (3.21) and (3.22) are the correct +starting point for all g and n. In particular, the c ghosts are not paired with any vertex +(a condition often assumed or presented as mandatory). This fact will help resolve some +difficulties for the 2-point function on the sphere. +Remember that there is no CKV and no c zero-mode for g ≥ 2. For the sphere g = 0 +and the torus g = 1, there are CKVs, indicating that there is a residual symmetry in (3.21) +and (3.22), which is the global conformal group of the worldsheet. It can be gauge fixed by +imposing conditions on the vertex operators.2 The simplest gauge fixing condition amounts +to fix the positions of Kc +g vertex operators through the Faddeev–Popov trick: +1 = ∆(σ0 +j ) +� +dξ +Kc +g +� +j=1 +δ(2)(σj − σ0(ξ) +j +), +σ0(ξ) +j += σ0 +j + δξσ0 +j , +δξσ0 +j = ξ(σ0 +j ), +(3.23) +where ξ is a conformal Killing vector, and the variation of σ was given in (2.7). We find +that +∆(σ0 +j ) = det ψi(σ0 +j ). +(3.24) +A priori, the positions σ0 +j are not the same as the one appearing in (2.163) (since both sets +are arbitrary): however, considering the same positions allows to cancel the factor (3.24) +with the same one in (2.163). +2In fact, it is only important to gauge fix for the sphere because the volume of the group is infinite. On +the other hand, the volume of the CKV group for the torus is finite-dimensional such that dividing by Ωckv +is not ambiguous. +62 + +Computation – Equation (3.24) +The first step is to compute ∆ in (3.23). For this, we decompose the CKV ξ on the +basis (2.104) +ξ(σ0 +j ) = αiψi(σ0 +j ) +and write the Gaussian integral: +1 = +� +Kc +g +� +j=1 +d2δσj e +−� +j(δσj,δσj) = ∆ +� +Kg +� +j=1 +dαi e +−� +j,i,i′(αiψi(σj),αi′ψi′(σj)) += ∆ +� +det ψi(σj) +�−1. +Again, we have reduced rigour in order to simplify the manipulations. +After inserting the identity (3.23) into (3.22), one can integrate over Kc +g vertex operator +positions to remove the delta functions – at the condition that there are at least Kc +g operators. +As a consequence, we learn that the proposed gauge fixing works only for n ≥ 1 if g = 1 or +n ≥ 3 if g = 0. This condition is equivalent to +χg,n = 2 − 2g − n < 0. +(3.25) +In this case, the factors det ψi(σ0 +j ) cancel and (3.21) becomes +Ag,n({ki}){αi} = g−χg,n +s +� +Mg +dMgt +� +dˆgb dˆgc +Kc +g +� +j=1 +ϵab +2 ca(σ0 +j )cb(σ0 +j ) +Mg +� +i=1 +(ˆµi, b)ˆg e−Sgh[ˆg,b,c] +× +� +n +� +i=Kc +g+1 +d2σi +� +ˆg +� Kc +g +� +j=1 +ˆVαj(kj; σ0 +j ) +n +� +i=Kcg+1 +ˆVαi(ki; σi) +� +m,ˆg +. +(3.26) +The result may be divided by a symmetry factor if the delta functions have solutions for +several points [193, sec. 5.3]. Performing the gauge fixing for the other cases (in particular, +g = 0, n = 2 and g = 1, n = 0) is more subtle (Section 3.1.3 and [193]). +The amplitude can be rewritten in two different ways. First, the ghost insertions can be +rewritten in terms of a ghost correlation functions +Ag,n({ki}){αi} = g−χg,n +s +� +Mg +dMgt +� +n +� +i=Kc +g+1 +d2σi +� +ˆg +� Kc +g +� +j=1 +ϵab +2 ca(σ0 +j )cb(σ0 +j ) +Mg +� +i=1 +(ˆµi, b)ˆg +� +gh,ˆg +× +� Kc +g +� +j=1 +ˆVαj(kj; σ0 +j ) +n +� +i=Kc +g+1 +ˆVαi(ki; σi) +� +m,ˆg +. +(3.27) +This form is particularly interesting because it shows that, before integration over the mod- +uli, the amplitudes factorize. This is one of the main advantage of the conformal gauge, since +the original complicated amplitude (3.6) for a QFT on a dynamical spacetime reduces to +the product of two correlation functions of QFTs on a fixed curved background. In fact, the +situation is even simpler when taking a flat background ˆg = δ since both the ghost and mat- +ter sectors are CFTs and one can employ all the tools from two-dimensional CFT (Part I) to +perform the computations and mostly forget about the path integral origin of these formulas. +This approach is particularly fruitful for off-shell (Chapter 11) and superstring amplitudes +(Chapter 17). +63 + +Remark 3.2 (Amplitudes in 2d gravity) The derivation of amplitudes for 2d gravity +follows the same procedure, up to two differences: 1) there is an additional decoupled (before +moduli and position integrations) gravitational sector described by the Liouville field, 2) the +matter and gravitational action are not CFTs if the original matter was not. +A second formula can be obtained by bringing the c-ghost on top of the matter vertex +operators which are at the same positions +Ag,n({ki}){αi} = g−χg,n +s +� +Mg +dMgt +� +n +� +i=Kc +g+1 +d2σi +� +ˆg +� Mg +� +i=1 +ˆBi +Kc +g +� +j=1 +ˆ +Vαj(kj; σ0 +j ) +n +� +i=Kc +g+1 +ˆVαi(ki; σi) +� +ˆg +, +(3.28) +and where +ˆ +Vαj(kj; σ0 +j ) := ϵab +2 ca(σ0 +j )cb(σ0 +j ) ˆVαj(kj; σ0 +j ), +ˆBi := (ˆµi, b)ˆg. +(3.29) +The operators Vαi(ki; σ0 +j ) (a priori off-shell) are called unintegrated operators, by opposition +to the integrated operators Vαi(ki). We will see that both are natural elements of the BRST +cohomology. +To stress that the ˆBi insertions are really an element of the measure, it is finally possible +to rewrite the previous expression as +Ag,n({ki}){αi} = g−χg,n +s +� +Mg×Cn−Kcg +� Mg +� +i=1 +ˆBi dti +Kc +g +� +j=1 +ˆ +Vαi(ki; σ0 +j ) +n +� +i=Kc +g+1 +ˆVαi(ki; σi) d2σi +� +ˆg +� +ˆg +. +(3.30) +The result (3.28) suggests a last possibility for improving the expression of the amplitude. +Indeed, the different vertex operators don’t appear symmetrically: some are integrated over +and other come with c ghosts. Similarly, the two types of integrals have different roles: the +moduli are related to geometry while the positions look like external data (vertex operators). +However, punctures can obviously be interpreted as part of the geometry, and one may +wonder if it is possible to unify the moduli and positions integrals. It is, in fact, possible to +put all vertex operators and integrals on the same footing by considering the amplitude to +be defined on the moduli space Mg,n of genus-g Riemann surfaces with n punctures instead +of just Mg [193] (see also Section 11.3.1). +3.1.3 +Gauge fixing: 2-point amplitude +As discussed at the end of Section 3.1.1, it has long been believed that the tree-level 2-point +amplitude vanishes. There were two main arguments: there are not sufficiently many vertex +operators 1) to fix completely the SL(2, C) invariance or 2) to saturate the number of c- +ghost zero-modes. Let’s review both points and then explain why they are incorrect. We +will provide the simplest arguments, referring the reader to the literature [66, 208] for more +general approaches. +For simplicity, we consider the flat metric ˆg = δ and an orthonormal basis of CKV. The +two weight-(1, 1) matter vertex operators are denoted as Vk(z, ¯z) and Vk′(z′, ¯z′) such that +the 2-point correlation function on the sphere reads (see Chapters 6 and 7 for more details): +⟨Vk(z, ¯z)Vk′(z′, ¯z′)⟩S2 = i (2π)Dδ(D)(k + k′) +|z − z′|4 +. +(3.31) +The numerator comes from the zero-modes ei(k+k′)·x for a target spacetime with a Lorentzian +signature [48, p. 866, 193] (required to make use of the on-shell condition). +64 + +Review of the problem +The tree-level amplitude (3.20) for n = 2 reads: +A0,2(k, k′) = +CS2 +Vol K0,0 +� +d2zd2z′ ⟨Vk(z, ¯z)Vk′(z′, ¯z′)⟩S2 , +(3.32) +where K0,n is the CKV group of Σ0,n, the sphere with n punctures. In particular, the group +of the sphere without puncture is K0,0 = PSL(2, C). The normalization of the amplitude is +CS2 = 8πα′−1 for gs = 1 [193, 249]. Since there are two insertions, the symmetry can be +partially gauge fixed by fixing the positions of the two punctures to z = 0 and z′ = ∞. In +this case, the amplitude (3.32) becomes: +A0,2(k, k′) = +CS2 +Vol K0,2 +⟨Vk(∞, ∞)Vk′(0, 0)⟩S2 , +(3.33) +where K0,2 = R∗ ++×U(1) is the CKV group of the 2-punctured sphere – containing dilatations +and rotations.3 Since the volume of this group is infinite Vol K0,2 = ∞, it looks like A0,2 = 0. +However, this forgets that the 2-point correlation function (3.31) contains a D-dimensional +delta function. +The on-shell condition implies that the conservation of the momentum +k + k′ = 0 is automatic for one component, such that the numerator in (3.33) contains a +divergent factor δ(0): +A0,2(k, k′) = (2π)D−1δ(D−1)(k + k′) CS2 2πi δ(0) +Vol K0,2 +. +(3.34) +Hence, (3.33) is of the form A0,2 = ∞/∞ and one should be careful when evaluating it. +The second argument relies on a loophole in the understanding of the gauge fixed amp- +litude (3.28). The result (3.28) is often summarized by saying that one can go from (3.20) +to (3.28) by replacing Kc +g integrated vertices +� +V by unintegrated vertices c¯cV in order to +saturate the ghost zero-modes and to obtain a non-zero result. For g = 0, this requires 3 +unintegrated vertices. But, since there are only two operators in (3.32), this is impossible +and the result must be zero. However, this is also incorrect because it is always possible +to insert 6 c zero-modes, as show the formulas (2.163) and (3.27). Indeed, they are part +of how the path integral measure is defined and do not care of the matter operators. The +question is whether they can be attached to vertex operators (for aesthetic reasons or more +pragmatically to get natural states of the BRST cohomology). To find the correct result +with ghosts requires to start with (3.27) and to see how this can be simplified when there +are only two operators. +Computation of the amplitude +In this section, we compute the 2-point amplitude from (3.33): +A0,2(k, k′) = +CS2 +Vol K0,2 +⟨Vk(∞, ∞)Vk′(0, 0)⟩S2 . +(3.35) +The volume of K0,2 reads (by writing a measure invariant under rotations and dilatations, +but not translations nor special conformal transformations) [53, 60]: +Vol K0,2 = +� d2z +|z|2 = 2 +� 2π +0 +dσ +� ∞ +0 +dr +r , +(3.36) +3The subgroup and the associated measure depend on the locations of the two punctures. +65 + +by doing the change of variables z = reiσ. Since the volume is infinite, it must be regu- +larized. A first possibility is to cut-off a small circle of radius ϵ around r = 0 and r = ∞ +(corresponding to removing the two punctures at z = 0, ∞). A second possibility consists +in performing the change of variables r = eτ and to add an imaginary exponential: +Vol K0,2 = 4π +� ∞ +0 +dr +r = 4π +� ∞ +−∞ +dτ = 4π lim +ε→0 +� ∞ +−∞ +dτ eiετ = 4π × 2π lim +ε→0 δ(ε), +(3.37) +such that the regularized volume reads +Volε K0,2 = 8π2 δ(ε). +(3.38) +In fact, τ can be interpreted as the Euclidean worldsheet time on the cylinder since r +corresponds to the radial direction of the complex plane. +Since the worldsheet is an embedding into the target spacetime, both must have the +same signature. As a consequence, for the worldsheet to be also Lorentzian, the formula +(3.37) must be analytically continued as ε = −iE and τ = it such that +VolM,E K0,2 = 8π2i δ(E), +(3.39) +where the subscript M reminds that one considers the Lorentzian signature. Inserting this +expression in (3.34) and taking the limit E → 0, it looks like the two δ(0) will cancel. +However, we need to be careful about the dimensions. Indeed, the worldsheet time τ and +energy E are dimensionless, while the spacetime time and energy are not. Thus, it is not +quite correct to cancel directly both δ(0) since they don’t have the same dimensions. In order +to find the correct relation between the integrals in (3.37) and of the zero-mode in (3.31), +we can look at the mode expansion for the scalar field (removing the useless oscillators): +X0(z, ¯z) = x0 + i +2 α′k0 ln |z|2 = x0 + iα′k0τ, +(3.40) +where the second equality follows by setting z = eτ. After analytic continuation k0 = −ik0 +M, +X0 = iX0 +M, x0 = ix0 +M and τ = it, we find [265, p. 186]: +X0 +M = x0 +M + α′k0 +Mt. +(3.41) +This indicates that the measure of the worldsheet time in (3.39) must be rescaled by 1/α′k0 +M +such that: +VolM K0,2 −→ 8π2i δ(0) +α′k0 +M += CS2 2πi δ(0) +2k0 +M +. +(3.42) +This is equivalent to rescale E by α′k0 and to use δ(ax) = a−1δ(x). +Ultimately, the 2-point amplitude becomes (removing the subscript on k0): +A0,2(k, k′) = 2k0(2π)D−1δ(D−1)(k + k′) +(3.43) +and matches the QFT formula (3.16). We see that taking into account the scale of the +coordinates is important to reproduce this result. +The computation displayed here presents some ambiguities because of the regularization. +However, this ambiguity can be fixed from unitarity of the scattering amplitudes. A more +general version of the Faddeev–Popov gauge fixing has been introduced in [66] to avoid +dealing altogether with infinities. +It is an interesting question whether these techniques +can be extended to the compute the tree-level 1- and 0-point amplitudes on the sphere. In +most cases, the 1-point amplitude is expected to vanish since 1-point correlation functions +66 + +of primary operators other than the identity vanish in unitary CFTs.4 The 0-point function +corresponds to the sphere partition function: the saddle point approximation to leading +order allows to relate it to the spacetime action evaluated on the classical solution φ0, +Z0 ∼ e−S[φ0]/ℏ. Since the normalization is not known and because S[φ0] is expected to +be infinite, only comparison between two spacetimes should be meaningful (à la Gibbons– +Hawking–York [190, sec. 4.1]). In particular, for Minkowski spacetime we find naively +Z0 ∼ δ(D)(0) +Vol K0 +, +(3.44) +which is not well-defined. This question has no yet been investigated. +Expression with ghosts +There are different ways to rewrite the 2-point amplitude in terms of ghosts. In all cases, one +correctly finds the 6 insertions necessary to get a non-vanishing result since, by definition, +it is always possible to rewrite the Faddeev–Popov determinant in terms of ghosts. A first +approach is to insert 1 = +� +d2z δ(2)(z) inside (3.32) to mimic the presence of a third operator. +This is equivalent to use the identity +⟨0| c−1¯c−1c0¯c0c1¯c1 |0⟩ = 1 +(3.45) +inside (3.33), leading to: +A0,2(k, k′) = +CS2 +Vol K0,2 +⟨Vk(∞, ∞)c0¯c0 Vk′(0, 0)⟩S2 , +(3.46) +where Vk(z, ¯z) = c¯cVk(z, ¯z). This shows that (3.16) can also be recovered using the correct +insertions of ghosts. The presence of c0¯c0 can be expected from string field theory since they +appear in the kinetic term (10.115). +The disadvantage of this formula is to still contain the infinite volume of the dilatation +group. It is also possible to introduce ghosts for the more general gauge fixing presented +in [66]. An alternative approach has been proposed in [208]. +3.2 +BRST quantization +The symmetries of a Lagrangian dictate the possible terms which can be considered. This +continues to hold at the quantum level and the counter-terms introduced by renormalization +are constrained by the symmetries. However, if the path integral is gauge fixed, the original +symmetry is no more available for this purpose. Fortunately, one can show that there is +a global symmetry (with anticommuting parameters) remnant of the local symmetry: the +BRST symmetry. It ensures consistency of the quantum theory. It also provides a direct +access to the physical spectrum. +The goal of this section is to provide a general idea of the BRST quantization for the +worldsheet path integral. A more detailed CFT analysis and the consequence for string +theory are given in Chapter 8. The reader is assumed to have some familiarity with the +BRST quantization in field theory – a summary is given in Appendix C.2. +4The integral over the zero-mode gives a factor δ(D)(k) which implies k = 0. +At zero momentum, +the time scalar X0 is effectively described by unitary CFT. However, there can be some subtleties when +considering marginal operator. +67 + +3.2.1 +BRST symmetry +The partition function (2.159) is not the most suitable to display the BRST symmetry. +The first step is to restore the dependence in the original metric gab by introducing a delta +function +Zg = +� +Mg +dMgt +Ωckv[g] +� +dggab dgΨ dgb d′ +gc δ +�√ggab − +� +ˆgˆgab +� Mg +� +i=1 +(φi, b)g e−Sm[g,Ψ]−Sgh[g,b,c]. +(3.47) +Note that it is necessary to use the traceless gauge fixing condition (2.152) as it will become +clear. The delta function is Fourier transformed in an exponential thanks to an auxiliary +bosonic field: +Zg = +� +Mg +dMgt +Ωckv[g] +� +dggab dgBab dgΨ dgb d′ +gc +Mg +� +i=1 +(φi, b)g e−Sm[g,Ψ]−Sgf[g,ˆg,B]−Sgh[g,b,c] +(3.48) +where the gauge-fixing action reads: +Sgf[g, ˆg, B] = − i +4π +� +d2σ Bab�√ggab − +� +ˆgˆgab +� +. +(3.49) +Varying the action with respect to the auxiliary field Bab, called the Nakanish–Lautrup field, +produces the gauge-fixing condition. +The BRST transformations are +δϵgab = iϵ Lcgab, +δϵΨ = iϵ LcΨ, +δϵca = iϵ Lcca, +δϵbab = ϵ Bab, +δϵBab = 0, +(3.50) +where ϵ is a Grassmann parameter (anticommuting number) independent of the position. +If the traceless gauge fixing (2.152) is not used, then Bab is not traceless: in that case, the +variation δϵbab will generate a trace, which is not consistent. Since the transformations act +on the matter action Sm as a diffeomorphism with vector ϵca, it is obvious that it is invariant +by itself. It is easy to show that the transformations (3.50) leave the total action invariant +in (3.48). The invariance of the measure is given in [193]. +Remark 3.3 (BRST transformations with Weyl ghost) One can also consider the ac- +tion (2.153) with the Weyl ghost. +In this case, the transformation law of the metric is +modified and the Weyl ghost transforms as a scalar: +δϵgab = iϵ Lcgab + iϵ gabcw, +δϵcw = iϵ Lccw. +(3.51) +The second term in δϵgab is a Weyl transformation with parameter ϵcw. Moreover, bab and +Bab are not symmetric traceless. +The equation of motion for the auxiliary field is +Bab = i Tab := i +� +T m +ab + T gh +ab +� +, +(3.52) +where the RHS is the total energy–momentum tensor (matter plus ghosts). Integrating it +out imposes the gauge condition gab = ˆgab and yields the modified BRST transformations +δϵΨ = iϵ LcΨ, +δϵca = iϵ Lcca, +δϵbab = iϵ Tab. +(3.53) +Without starting with the path integral (3.48) with auxiliary field, it would have been +difficult to guess the transformation of the b ghost. Since ca is a vector, one can also write +δϵca = ϵ cb∂bca. +(3.54) +68 + +Associated to this symmetry is the BRST current ja +B and the associated conserved BRST +charge QB +QB = +� +dσ j0 +B. +(3.55) +The charge is nilpotent +Q2 +B = 0, +(3.56) +and, through the presence of the c-ghost in the BRST transformation, the BRST charge has +ghost number one +Ngh(QB) = 1. +(3.57) +Variations of the matter fields can be written as +δϵΨ = i [ϵQB, Ψ]±. +(3.58) +Note that the energy–momentum tensor is BRST exact +Tab = [QB, bab]. +(3.59) +3.2.2 +BRST cohomology and physical states +Physical state |ψ⟩ are elements of the absolute cohomology of the BRST operator: +|ψ⟩ ∈ H(QB) := ker QB +Im QB +, +(3.60) +or, more explicitly, closed but non-exact states: +QB |ψ⟩ = 0, +∄ |χ⟩ : |ψ⟩ = QB |χ⟩ . +(3.61) +The adjective “absolute” is used to distinguish it from two other cohomologies (relative +and semi-relative) defined below. Two states of the cohomology differing by an exact state +represent identical physical states: +|ψ⟩ ∼ |ψ⟩ + QB |Λ⟩ . +(3.62) +This equivalence relation, translated in terms of spacetime fields, correspond to spacetime +gauge transformations. In particular, it contains the (linearized) reparametrization invari- +ance of the spacetime metric in the closed string sector, and, for the open string sector, it +contains Yang–Mills symmetries. We will find that it corresponds to the gauge invariance +of free string field theory (Chapter 10). +However, physical states satisfy two additional constraints (remember that bab is traceless +symmetric): +� +dσ bab |ψ⟩ = 0. +(3.63) +These conditions are central to string (field) theory, so they will appear regularly in this +book. For this reason, it is useful to provide first some general motivations, and to refine +the analysis later since the CFT language will be more appropriate. Moreover, these two +conditions will naturally emerge in string field theory. +In order to introduce some additional terminology, let’s define the following quantities:5 +b+ := +� +dσ b00, +b− := +� +dσ b01. +(3.64) +5The objects b± are zero-modes of the b ghost fields. They correspond (up to a possible irrelevant factor) +to the modes b± +0 in the CFT formulation of the ghost system (7.132). +69 + +The semi-relative and relative cohomologies H−(QB) and H0(QB) are defined as6 +H−(QB) = H(QB) ∩ ker b−, +H0(QB) = H−(QB) ∩ ker b+. +(3.65) +The first constraint arises as a consequence of the topology of the closed string worldsheet: +the spatial direction is a circle, which implies that the theory must be invariant under +translations along the σ direction (the circle is invariant under rotation). However, choosing +a parametrization implies to fix an origin for the spatial direction: this is equivalent to +a gauge fixing condition. +As usual, this implies that the corresponding generator Pσ of +worldsheet spatial translations (2.26) must annihilate the states: +Pσ |ψ⟩ = 0. +(3.66) +This is called the level-matching condition. Using (3.59), this can be rewritten as +Pσ |ψ⟩ = +� +dσ T01 |ψ⟩ = +� +dσ {QB, b01} |ψ⟩ = QB +� +dσ b01 |ψ⟩ , +(3.67) +since QB |ψ⟩ = 0 for a state |ψ⟩ in the cohomology. +The simplest way to enforce this +condition is to set the state on which QB acts to zero:7 +b− |ψ⟩ = 0, +(3.68) +which is equivalent to one of the conditions in (3.63). +The second condition does not follow as simply. The Hilbert space can be decomposed +according to b+ as +H− := H↓ ⊕ H↑, +H↓ := H0 := H− ∩ ker b+. +(3.69) +Indeed, b+ is a Grassmann variable and generates a 2-state system. In the ghost sector, the +two Hilbert spaces are generated from the ghost vacua | ↓⟩ and | ↑⟩ obeying +b+ | ↓⟩ = 0, +b+ | ↑⟩ = | ↓⟩ . +(3.70) +The action of the BRST charge on states |ψ↓⟩ ∈ H↓ and |ψ↑⟩ ∈ H↑ follow from these relations +and from the commutation relation (3.59): +QB |ψ↓⟩ = H |ψ↑⟩ , +QB |ψ↑⟩ = 0, +(3.71) +where H is the worldsheet Hamiltonian defined in (2.26). To prove this relation, start first +with H |ψ↑⟩, then use (3.59)) to get the LHS of the first condition; then apply QB to get +the second condition (using that QB commutes with H, and b+ with any other operators +building the states). For H ̸= 0, the state |ψ↓⟩ is not in the cohomology and |ψ↑⟩ is exact. +Thus, the exact and closed states are +Im QB = +� +|ψ↑⟩ ∈ H↑ | H |ψ↑⟩ ̸= 0 +� +, +(3.72a) +ker QB = +� +|ψ↑⟩ ∈ H↑ +� +∪ +� +|ψ↓⟩ ∈ H↓ | H |ψ↓⟩ = 0 +� +. +(3.72b) +This implies that eigenstates of H in the cohomology satisfy the on-shell condition: +H |ψ⟩ = 0. +(3.73) +6The BRST cohomologies described in this section are slightly different from the ones used in the rest +of this book. +To distinguish them, indices are written as superscripts in this section, and as subscripts +otherwise. +7The reverse is not true. We will see in Section 3.2.2 the relation between the two conditions in more +details. +70 + +This is consistent with the fact that scattering amplitudes involve on-shell states. In this +case, |ψ↑⟩ is not exact and is thus a member of the cohomology H(QB), as well as |ψ↓⟩ since +it becomes close. But, the Hilbert space H↑ must be rejected for two reasons: there would +be an apparent doubling of states and scattering amplitudes would behave badly. The first +problem arises because one can show that the cohomological subspaces of each space are +isomorphic: H↓(QB) ≃ H↑(QB). Hence, keeping both subspaces would lead to a doubling +of the physical states. For the second problem, consider an amplitude where one of the +external state is built from |ψ↑⟩: the amplitude vanishes if the states are off-shell since the +state |ψ↑⟩ is exact, but it does not vanish on-shell [193, ch. 4]. This means that it must +be proportional to δ(H). But, general properties in QFT forbid such dependence in the +amplitude (only poles and cuts are allowed, except if D = 2). Projecting out the states in +H↑ is equivalent to require +b+ |ψ⟩ = 0 +(3.74) +for physical states, which is the second condition in (3.63). +In fact, this condition can be obtained very similarly as the b− = 0 condition: using the +expression of H (2.26) and the commutation relation (3.59), (3.73) is equivalent to +QB +� +dσ b00 |ψ⟩ = 0. +(3.75) +Hence, imposing (3.74) allows to automatically ensure that (3.73) holds. +Since the on-shell character (3.73) of the BRST states and of the BRST symmetry are +intimately related to the construction of the worldsheet integral, one can expect difficulty +for going off-shell. +3.3 +Summary +In this chapter, we derived general formulas for string scattering amplitudes. The general +BRST formalism has been summarized. Moreover, we gave general motivations for restrict- +ing the absolute cohomology to the smaller relative cohomology. +In Chapter 8, a more +precise derivation of the BRST cohomology is worked out. It includes also a proof of the +no-ghost theorem: the ghosts and the negative norm states (in Minkowski signature) are +unphysical particles and should not be part of the physical states. This theorem asserts that +it is indeed the case. It will also be the occasion to recover the details of the spectrum in +various cases. +3.4 +Suggested readings +• The delta function approach to the gauge fixing is described in [193, sec. 3.3, 151, +sec. 15.3.2], with a more direct computation is in [128]. +• The most complete references for scattering amplitudes in the path integral formalism +are [53, 193]. +• Computation of the tree-level 2-point amplitude [66, 208] (for discussions of 2-point +function, see [53, p. 936–7, 207, 60, 61, 48, p. 863–4]). +• The BRST quantization of string theory is discussed in [155, 39, 193, chap. 4]. For +a general discussion see [105, 247, 251]. The use of an auxiliary field is considered +in [252, sec. 3.2]. +71 + +Chapter 4 +Worldsheet path integral: +complex coordinates +Abstract +In the two previous chapters, the amplitudes computed from the worldsheet +path integrals have been written covariantly for a generic curved background metric. In +this chapter, we start to use complex coordinates and finally take the background metric to +be flat. This is the usual starting point for computing amplitudes since it allows to make +contact with CFTs and to employ tools from complex analysis. We first recall few facts on +2d complex manifolds before briefly describing how to rewrite the scattering amplitudes in +complex coordinates. +4.1 +Geometry of complex manifolds +Choosing a flat background metric simplifies the computations. However, we have seen in +Section 2.3 that there is a topological obstruction to get a globally flat metric. The solution +is to work with coordinate patches (σ0, σ1) = (τ, σ) such that the background metric ˆgab is +flat in each patch (conformally flat gauge): +ds2 = gabdσadσb = e2φ(τ,σ)� +dτ 2 + dσ2� +, +(4.1) +or +gab = e2φδab, +ˆgab = δab. +(4.2) +To simplify the notations, we remove the dependence in the flat metric and the hat for +quantities (like the vertex operators) expressed in the background metric when no confusion +is possible. +Introducing complex coordinates +z = τ + iσ, +¯z = τ − iσ, +(4.3a) +τ = z + ¯z +2 +, +σ = z − ¯z +2i +, +(4.3b) +the metric reads1 +ds2 = 2gz¯zdzd¯z = e2φ(z,¯z)|dz|2. +(4.4) +1In Section 6.1, we provide more details on the relation between the worldsheet (viewed as a cylinder or +a sphere) and the complex plane. +72 + +The metric and its inverse can also be written in components: +gz¯z = e2φ +2 , +gzz = g¯z¯z = 0, +(4.5a) +gz¯z = 2e−2φ, +gzz = g¯z¯z = 0. +(4.5b) +Equivalently, the non-zero components of the background metric are +ˆgz¯z = 1 +2, +ˆgz¯z = 2. +(4.6) +An oriented two-dimensional manifold is a complex manifold: this means that there exists +a complex structure, such that the transition functions and changes of coordinates between +different patches are holomorphic at the intersection of the two patches: +w = w(z), +¯w = ¯w(¯z). +(4.7) +For such a transformation, the Liouville mode transforms as +e2φ(z,¯z) = +���� +∂w +∂z +���� +2 +e2φ(w, ¯ +w) +(4.8) +such that +ds2 = e2φ(w, ¯ +w)|dw|2. +(4.9) +This shows also that a conformal structure (2.12) induces a complex structure since the +transformation law of φ is equivalent to a Weyl rescaling. +The integration measures are related as +d2σ := dτdσ = 1 +2 d2z, +d2z := dzd¯z. +(4.10) +Due to the factor of 2 in the expression, the delta function δ(2)(z) also gets a factor of 2 +with respect to δ(2)(σ) +δ(2)(z) = 1 +2 δ(2)(σ). +(4.11) +Then, one can check that +� +d2z δ(2)(z) = +� +d2σ δ(2)(σ) = 1. +(4.12) +The basis vectors (derivatives) and one-forms can be found using the chain rule: +∂z = 1 +2 (∂τ − i∂σ), +∂¯z = 1 +2 (∂τ + i∂σ), +(4.13a) +dz = dτ + idσ, +d¯z = dτ − idσ. +(4.13b) +The Levi–Civita (completely antisymmetric) tensor is normalized by +ϵ01 = ϵ01 = 1. +(4.14a) +ϵz¯z = i +2, +ϵz¯z = −2i, +(4.14b) +remembering that it transforms as a density. Integer indices run over local frame coordinates. +The different tensors can be found from the tensor transformation law. For example, the +components of a vector V a in both systems are related by +V z = V 0 + iV 1, +V ¯z = V 0 − iV 1 +(4.15) +73 + +such that +V = V 0∂0 + V 1∂1 = V z∂z + V ¯z∂¯z. +(4.16) +For holomorphic coordinate transformations (4.7), the components of the vector do not mix: +V w = ∂w +∂z V z, +V ¯ +w = ∂ ¯w +∂¯z V ¯z. +(4.17) +This implies that the tangent space of the Riemann surface is decomposed into holomorphic +and anti-holomorphic vectors:2 +TΣg ≃ TΣ+ +g ⊕ TΣ− +g , +(4.18a) +V z∂z ∈ TΣ+ +g , +V ¯z∂¯z ∈ TΣ− +g , +(4.18b) +as a consequence of the existence of a complex structure. Similarly, the components of a +1-form ω – which is the only non-trivial form on Σg – can be written in terms of the real +coordinates as: +ωz = 1 +2 (ω0 − iω1), +ω¯z = 1 +2 (ω0 + iω1) +(4.19) +such that +ω = ω0dσ0 + ω1dσ1 = ωzdz + ω¯zd¯z. +(4.20) +Hence, a 1-form is decomposed into complex (1, 0)- and (0, 1)-forms: +T ∗Σg ≃ Ω1,0(Σg) ⊕ Ω0,1(Σg), +(4.21a) +ωzdz ∈ Ω1,0(Σg), +ω¯zd¯z ∈ Ω0,1(Σg), +(4.21b) +since both components will not mixed under holomorphic changes of coordinates (4.7). Fi- +nally, the metric provides an isomorphism between TΣ+ +g and Ω0,1(Σg), and between TΣ− +g +and Ω1,0(Σg), since it can be used to lower/raise an index while converting it from holo- +morphic to anti-holomorphic, or conversely: +Vz = gz¯zV ¯z, +V¯z = gz¯zV z. +(4.22) +This can be generalized further by considering components with more indices: all anti- +holomorphic indices can be converted to holomorphic indices thanks to the metric: +T +q++p− +���� +z···z +z···z +���� +p++q− += (gz¯z)p−(gz¯z)q−T +q+ +���� +z···z +q− +���� +¯z···¯z +z···z +���� +p+ +¯z···¯z +���� +p− +. +(4.23) +Hence, it is sufficient to study (p, q)-tensors with p upper and q lower holomorphic indices. +In this case, the transformation rule under (4.7) reads +T +q +���� +w···w +w···w +���� +p += +�∂w +∂z +�n +T +q +���� +z···z +z···z +���� +p +, +n := q − p. +(4.24) +The number n ∈ Z is called the helicity or rank.3 The set of helicity-n tensors is denoted +by T n. +The first example is vectors (or equivalently 1-forms): V z ∈ T 1, Vz ∈ T −1. The second +most useful case is traceless symmetric tensors, which are elements of T ±2. Consider a +2However, at this stage, each component can still depend on both z and ¯z: V z = V z(z, ¯z) and V ¯z = +V ¯z(z, ¯z). +3In fact, it is even possible to consider n ∈ Z + 1/2 to describe spinors. +74 + +traceless symmetric tensor T ab = T ba and gabT ab = 0: this implies T 01 = T 10 and T 00 = +−T 11 in real coordinates. The components in complex coordinates are: +T zz = 2(T 00 + iT 01) ∈ T 2, +T ¯z¯z = 2(T 00 − iT 01) ∈ T −2, +T z +z = 0. +(4.25) +Note that +Tzz = gz¯zgz¯zT ¯z¯z = 1 +2(T 00 − iT 01), +(4.26) +and T z +z = gz¯zT z¯z ∈ T 0 corresponds to the trace. +Computation – Equation (4.25) +T zz = +�∂z +∂τ +�2 +T 00 + +� ∂z +∂σ +�2 +T 11 + 2 ∂z +∂τ +∂z +∂σ T 01 = T 00 − T 11 + 2i T 01. +Stokes’ theorem in complex coordinates follows directly from (B.10): +� +d2z (∂zvz + ∂¯zv¯z) = −i +� � +dz v¯z − d¯zvz� += −2i +� +∂R +(vzdz − v¯zd¯z), +(4.27) +where the integration contour is anti-clockwise. To obtain this formula, note that d2x = 1 +2d2z +and ϵz¯z = i/2, such that the factor 1/2 cancels between both sides. +4.2 +Complex representation of path integral +In the previous section, we have found that tensors of a given rank are naturally decomposed +into different subspaces thanks to the complex structure of the manifold. +Accordingly, +complex coordinates are natural and one can expect most objects in string theory to split +similarly into holomorphic and anti-holomorphic sectors (or left- and right-moving). This +will be particularly clear using the CFT language (Chapter 6). The main difficulty for this +program is due to the matter zero-modes. In this section, we focus on the path integral +measure and expression of the ghosts. +There is, however, a subtlety in displaying explicitly the factorization: the notion of +“holomorphicity” depends on the metric (because the complex structure must be compatible +with the metric for an Hermitian manifold). Since the metric depends on the moduli which +are integrated over in the path integral, it is not clear that there is a consistent holomorphic +factorization. We will not push the question of achieving a global factorization further (but +see Remark 4.1) to focus instead on the integrand. The latter is local (in moduli space) and +there is no ambiguity. +The results of the previous section indicate that the basis of Killing vectors (2.104) and +quadratic differentials (2.76) split into holomorphic and anti-holomorphic components: +ψi(z, ¯z) = ψz +i ∂z + ψ¯z +i ∂¯z, +φi(z, ¯z) = φi,zz(dz)2 + φi,¯z¯z(d¯z)2. +(4.28) +Similarly, the operators P1 (2.65a) and P † +1 (2.71) also split: +(P1ξ)zz = 2∇zξz = ∂zξ¯z, +(P1ξ)¯z¯z = 2∇¯zξ¯z = ∂¯zξz, +(4.29a) +(P † +1 T)z = −2∇zTzz = −4 ∂¯zTzz, +(P † +1 T)¯z = −2∇¯zT¯z¯z = −4 ∂zT¯z¯z +(4.29b) +for arbitrary vector ξ and traceless symmetric tensor T (in the background metric). As a +consequence, the components of Killing vectors and quadratic differentials are holomorphic +or anti-holomorphic as a function of z: +ψz = ψz(z), +ψ¯z = ψ¯z(¯z), +φzz = φzz(z), +φ¯z¯z = φ¯z¯z(¯z), +(4.30) +75 + +such that it makes sense to consider a complex basis instead of the previous real basis: +ker P1 = Span{ψK(z)} ⊕ Span{ ¯ψK(¯z)}, +K = 1, . . . , Kc +g, +(4.31a) +ker P † +1 = Span{φI(z)} ⊕ Span{¯φI(¯z)}, +I = 1, . . . , Mc +g. +(4.31b) +The last equation can inspire to search for a similar rewriting of the moduli parameters. +In fact, the moduli space itself is a complex manifold and can be endowed with complex +coordinates [173, 193]: +mI = t2I−1 + it2I, +¯mI = t2I−1 − it2I, +I = 1, . . . , Mc +g +(4.32) +with the integration measure +dMgt = d2Mc +gm. +(4.33) +The last ingredient to rewrite the vacuum amplitudes (2.136) is to obtain the determin- +ants. The inner-products of vector and traceless symmetric fields also factorize: +(T1, T2) = 2 +� +d2σ +� +ˆg ˆgacgbdT1,abT2,cd = 4 +� +d2z +� +T1,zzT2,¯z¯z + T1,¯z¯zT2,zz +� +, +(4.34a) +(ξ1, ξ2) = +� +d2σ +� +ˆg ˆgabξaξb = 1 +4 +� +d2z +� +ξz +1ξ¯z +2 + ξ¯z +1ξz +2 +� +. +(4.34b) +All inner-products are evaluated in the flat background metric. For (anti-)holomorphic fields, +only one term survives in each integral: since each field appears twice in the determinants +(φi, φj) and (φi, φj), the final expression is a square, which cancels against the squareroot +in (2.136). The remaining determinant involves the Beltrami differential (2.65b): +µizz = ∂i¯gzz, +µi¯z¯z = ∂i¯g¯z¯z +(4.35) +(¯gzz = 0 in our coordinates system, but its variation under a shift of moduli is not zero). +The basis can be changed to a complex basis such that the determinant of inner-products +between Beltrami and quadratic differentials is a modulus squared. All together, the different +formulas lead to the following rewriting of the vacuum amplitude : +Zg = +� +Mg +d2Mc +gm | det(φI, µJ)|2 +| det(φI, ¯φJ)| +det′ P † +1 P1 +| det(ψI, ¯ψJ)| +Zm[δ] +Ωckv[δ], +(4.36) +where the absolute values are to be understood with respect to the basis of P1 and P † +1 , for +example |f(mI)|2 := f(mI)f( ¯mI). +The same reasoning can be applied to the ghosts. The c and b ghosts are respectively +a vector and a symmetric traceless tensor, both with two independent components: it is +customary to define +c := cz, +¯c := c¯z, +b := bzz, +¯b := b¯z¯z. +(4.37) +In that case, the action (2.145) reads +Sgh[g, b, c] = 1 +2π +� +d2z +� +b∂¯zc + ¯b∂z¯c +� +. +(4.38) +The action is the sum of two holomorphic and anti-holomorphic contributions and it is +independent of φ(z, ¯z) as expected. In fact, the equations of motion are +∂zc = 0, +∂zb = 0, +∂¯z¯c = 0, +∂¯z¯b = 0, +(4.39) +76 + +such that b and c (resp. ¯b and ¯c) are holomorphic (anti-holomorphic) functions. Then, the +integration measure is simply +Mg +� +i=1 +Bi dti = +Mc +g +� +I=1 +BI ¯BI dmI ∧ ¯mI, +BI := (µI, b). +(4.40) +Note that BI does not contain ¯b(¯z), it is built only from b(z). +Finally, the vacuum amplitude (2.163) reads +Zg = +� +Mg +d2Mc +gm +Ωckv[δ]−1 +| det ψI(z0 +j )|2 +� +d(b,¯b) d(c, ¯c) +Kc +g +� +j=1 +c(z0 +j )¯c(¯z0 +j ) +Mc +g +� +I=1 +|(µI, b)|2 e−Sgh[b,c] Zm[δ]. +(4.41) +The c insertions are separated in holomorphic and anti-holomorphic components because, +at the end, only the zero-modes contribute. The measures are written as d(b,¯b) and d(c, ¯c) +because proving that they factorize is difficult (Remark 4.1). +Remark 4.1 (Holomorphic factorization) It was proven in [15, 27, 33] (see [173, sec. 9, +53, sec. VII, 237, sec. 3] for reviews) that the ghost and matter path integrals can be globally +factorized, up to a factor due to zero-modes. Such a result is suggested by the factorization +of the inner-products, which imply a factorization of the measures: the caveat is due to the +zero-mode determinants and matter measure. Interestingly, the factorization is possible only +in the critical dimension (2.125). +4.3 +Summary +In this chapter, we have introduced complex notations for the fields, path integral and +moduli space. +4.4 +Suggested readings +• Good references for this chapter are [24, 53, 172, 173, 193]. +• Geometry of complex manifolds is discussed in [24, sec. 6.2, 172, chap. 14, 53]. +77 + +Chapter 5 +Conformal symmetry in D +dimensions +Abstract +Starting with this chapter, we discuss general properties of conformal field the- +ories (CFT). The goal is not to be exhaustive, but to provide a short introduction and to +gather the concepts and formulas that are needed for string theory. However, the subject is +presented as a standalone topic such that it can be of interest for a more general public. +The conformal group in any dimension is introduced in this chapter. The specific case +D = 2, which is the most relevant for the current book, is developed in the following chapters. +5.1 +CFT on a general manifold +In this chapter and in the next one, we discuss CFTs as QFTs living on a spacetime M, +independently from string theory (there is no reference to a target spacetime). As such, we +will use spacetime notations together with some simplifications: coordinates are written as +xµ with µ = 0, . . . , D − 1 and time is written as x0 = t (x0 = τ) in Lorentzian (Euclidean) +signature. +Given a metric gµν on a D-dimensional manifold M, the conformal group CISO(M) is +the set of coordinate transformations (called conformal symmetries or isometries) +xµ −→ x′µ = x′µ(x) +(5.1) +which leaves the metric invariant up to an overall scaling factor: +gµν(x) −→ g′ +µν(x′) = ∂xρ +∂x′µ +∂xσ +∂x′ν gρσ(x) = Ω(x′)2gµν(x′). +(5.2) +This means that angles between two vectors u and v are left invariant under the transform- +ation: +u · v +|u| |v| = u′ · v′ +|u′| |v′|. +(5.3) +It is often convenient to parametrize the scale factor by an exponential +Ω := eω. +(5.4) +Considering an infinitesimal transformation +δxµ = ξµ, +(5.5) +78 + +the condition (5.2) becomes the conformal Killing equation +δgµν = Lξgµν = ∇µξν + ∇νξµ = 2 +d gµν∇ρξρ, +(5.6) +such that the scale factor is +Ω2 = 1 + 2 +d ∇ρξρ. +(5.7) +The vector fields ξ satisfying this equation are called conformal Killing vectors (CKV). Con- +formal transformations form a global subgroup of the diffeomorphism group: the generators +of the transformations do depend on the coordinates, but the parameters do not (for an +internal global symmetry, both the generators and the parameters don’t depend on the +coordinates). +The conformal group contains the isometry group ISO(M) of M as a subgroup, corres- +ponding to the case Ω = 1: +ISO(M) ⊂ CISO(M). +(5.8) +These transformations also preserve distances between points. The corresponding generators +of infinitesimal transformations are called Killing vectors and satisfies the Killing equation +δgµν = Lξgµν = ∇µξν + ∇νξµ = 0. +(5.9) +They form a subalgebra of the CKV algebra. +An important point is to be made for the relation between infinitesimal and finite trans- +formations: with spacetime symmetries it often happens that the first cannot be exponenti- +ated into the second. The reason is that the (conformal) Killing vectors may be defined only +locally, i.e. they are well-defined in a given domain but have singularities outside. When +this happens, they do not lead to an invertible transformation, which cannot be an element +of the group. These notions are sometimes confused in physics and the term of “group” is +used instead of “algebra”. We shall be careful in distinguishing both concepts. +Remark 5.1 (Isometries of M ⊂ Rp,q) In order to find the conformal isometries of a +manifold M which is a subset of Rp,q defined in (5.10), it is sufficient to restrict the trans- +formations of Rp,q to the subset M [206]. In the process, not all global transformations +generically survive. On the other hand, the algebra of local (infinitesimal) transformations +for M and Rp,q are identical since M is locally like Rp,q. +5.2 +CFT on Minkowski space +In this section, we consider the case where M = Rp,q (D = p + q) and where g = η is the +flat metric with signature (p, q): +η = diag(−1, . . . , −1 +� +�� +� +q +, 1, . . . , 1 +� �� � +p +). +(5.10) +The conformal Killing equation becomes +� +ηµν∆ + (D − 2)∂µ∂ν +� +∂ · ϵ = 0, +(5.11) +where ∆ is the D-dimensional Beltrami–Laplace operator for the metric ηµν. The case D = 2 +is relegated to the next chapter. For D > 2, one finds the following transformations: +translation: +ξµ = aµ, +(5.12a) +rotation & boost: +ξµ = ωµ +νxν, +(5.12b) +dilatation: +ξµ = λ xµ, +(5.12c) +SCT: +ξµ = bµx2 − 2b · x xµ, +(5.12d) +79 + +where ωµν is antisymmetric. The rotations include Lorentz transformations and SCT means +“special conformal transformation”. +All parameters {aµ, ωµν, λ, bµ} are constant. The generators are respectively denoted by +{Pµ, Jµν, D, Kµ}. The finite translations and rotations form the Poincaré group SO(p, q), +while the conformal group can be shown to be SO(p + 1, q + 1): +ISO(Rp,q) = SO(p, q), +CISO(Rp,q) = SO(p + 1, q + 1). +(5.13) +The dimension of this group is +dim SO(p + 1, q + 1) = 1 +2 (p + q + 2)(p + q + 1). +(5.14) +5.3 +Suggested readings +• References on higher-dimensional CFTs are [54, 196, 203, 206, 236]. +80 + +Chapter 6 +Conformal field theory on the +plane +Abstract +Starting with this chapter, we focus on two-dimensional Euclidean CFTs on the +complex plane (or equivalently the sphere). We start by describing the geometry of the +sphere and the relation to the complex plane and to the cylinder, in order to make contact +with the string worldsheet. Then, we discuss classical CFTs and the Witt algebra obtained +by classifying the conformal isometries of the complex plane. Then, we describe quantum +CFTs and introduce the operator formalism. This last section is the most important for this +book as it includes information on the operator product expansion, Hilbert space, Hermitian +and BPZ conjugations. +As described at the beginning of Chapter 5, we use spacetime notations for the coordin- +ates, but follow otherwise the normalization for the worldsheet. In particular, integrals are +normalized by 2π. However, the spatial coordinate on the cylinder is still written as σ to +avoid confusions: xµ = (τ, σ). +6.1 +The Riemann sphere +6.1.1 +Map to the complex plane +The Riemann sphere Σ0, which is diffeomorphic to the unit sphere S2, has genus g = 0 and +is thus the simplest Riemann surface. Its most straightforward description is obtained by +mapping it to the extended1 complex plane ¯C (also denoted ˆC), which is the complex plane +z ∈ C to which the point at infinity z = ∞ is added: +¯C = C ∪ {∞}. +(6.1) +One speaks about “the point at infinity” because all the points at infinity (i.e. the points z +such that |z| → ∞) +lim +r→∞ r eiθ := ∞ +(6.2) +are identified (the limit is independent of θ). +The identification can be understood by mapping (say) the south pole to the origin of +the plane and the north pole to infinity2 (Figure 6.1) through the stereographic projection +z = eiφ cot θ +2, +(6.3) +1This qualification will often be omitted. +2Note that the points are distinguished in order to write the map, but they have nothing special by +themselves (i.e. they are not punctures). +81 + +−−−−−−−−→ +Figure 6.1: Map from the Riemann sphere to the complex plane. The south and north poles +are denoted by the letter S and N, and the equatorial circle by E. +where (θ, φ) are angles on the sphere. Any circle on the sphere is mapped to a circle in the +complex plane. Conversely, the Riemann sphere can be viewed as a compactification of the +complex plane. +Introducing Cartesian coordinates (x, y) related to the complex coordinates by3 +z = x + iy, +¯z = x − iy, +(6.4a) +x = z + ¯z +2 +, +y = z − ¯z +2i +, +(6.4b) +the metric reads +ds2 = dx2 + dy2 = dzd¯z. +(6.5) +The relations between the derivatives in the two coordinate systems are easily found: +∂ := ∂z = 1 +2 (∂x − i∂y), +¯∂ := ∂¯z = 1 +2 (∂x + i∂y). +(6.6) +The indexed form will be used when there is a risk of confusion. If the index is omitted then +the derivative acts directly to the field next to it, for example +∂φ(z1)∂φ(z2) := ∂z1∂z2φ(z1)φ(z2). +(6.7) +Generically, the meromorphic and anti-meromorphic parts of a object will be denoted +without and with a bar, see (6.55) for an example. +The extended complex plane ¯C can be covered by two coordinate patches z ∈ C and +w ∈ C. In the first, the point at infinity (north pole) is removed, in the second, the origin +(south pole) is removed. On the overlap, the transition function is +w = 1 +z . +(6.8) +This description avoids to work with the infinity: studying the behaviour of f(z) at z = ∞ +is equivalent to study f(1/w) at w = 0. +Since any two-dimensional metric is locally conformally equivalent to the flat metric, it +is sufficient to work with this metric in each patch. This is particularly convenient for the +Riemann sphere since one patch covers it completely except for one point. +3General formulas can be found in Section 4.1 by replacing (τ, σ) with (x, y). In most cases, the conformal +factor is set to zero (φ = 0) in this chapter. +82 + +6.1.2 +Relation to the cylinder – string theory +The worldsheet of a closed string propagating in spacetime is locally topologically a cylinder +R × S1 of circumference L. In this section, we show that the cylinder can also be mapped +to the complex plane – and thus to the Riemann sphere – after removing two points. Since +the cylinder has a clear physical interpretation in string theory, it is useful to know how to +translate the results from the plane to the cylinder. +It makes also sense to define two-dimensional models on the cylinder independently of +a string theory interpretation since the compactification of the spatial direction from R to +S1 regulates the infrared divergences. Moreover, it leads to a natural definition of a “time” +and of an Hamiltonian on the Euclidean plane. +Denoting the worldsheet coordinates in Lorentzian signature by (t, σ) with4 +t ∈ R, +σ ∈ [0, L), +σ ∼ σ + L, +(6.9) +the metric reads +ds2 = −dt2 + dσ2 = −dσ+dσ−, +(6.10) +where the light-cone coordinates +dσ± = dt ± dσ +(6.11) +have been introduced. It is natural to perform a Wick rotation from the Lorentzian time t +to the Euclidean time +τ = it, +(6.12) +and the metric becomes +ds2 = dτ 2 + dσ2. +(6.13) +It is convenient to introduce the complex coordinates +w = τ + iσ, +¯w = τ − iσ +(6.14) +for which the metric is +ds2 = dwd ¯w. +(6.15) +Note that the relation to Lorentzian light-cone coordinates are +w = i(t + σ) = iσ+, +¯w = i(t − σ) = iσ−. +(6.16) +Hence, an (anti-)holomorphic function of w ( ¯w) depends only on σ+ (σ−) before the Wick +rotation: this leads to the identification of the left- and right-moving sectors with the holo- +morphic and anti-holomorphic sectors of the theory. +The cylinder can be mapped to the complex plane through +z = e2πw/L, +¯z = e2π ¯ +w/L, +(6.17) +and the corresponding metric is +ds2 = +� L +2π +�2 dzd¯z +|z|2 . +(6.18) +A conformal transformation brings this metric to the flat metric (6.5). The conventions +for the various coordinates and maps vary in the different textbooks. We have gathered in +Table A.1 the three main conventions and which references use which. +4Consistently with the comments at the beginning of Chapter 5, the Lorentzian worldsheet time is +denoted by t instead of τM. +83 + +−−−−−→ +−−−−−→ +Figure 6.2: Map from the cylinder to the sphere with two tubes, to the 2-punctured sphere +Σ0,2. +The map from the cylinder to the plane is found by sending the bottom end (corres- +ponding to the infinite past t → −∞) to the origin of the plane, and the top end (infinite +future t → ∞) to the infinity. Since the cylinder has two boundaries (its two ends) the map +excludes the point z = 0 and z = ∞ and one really obtains the space ¯C − {0, ∞} = C∗. +This space can, in turn, be mapped to the 2-punctured Riemann sphere Σ0,2. +The physical interpretation for the difference between Σ0 and Σ0,2 is simple: since one +considers the propagation of a string, it means that the worldsheet corresponds to an amp- +litude with two external states, which are the mapped to the sphere as punctures (Figure 6.2, +Section 3.1.1). Removing the external states (yielding the tree-level vacuum amplitude) cor- +responds to gluing half-sphere (caps) at each end of the cylinder (Figure 6.3). Then, it can +be mapped to the Riemann sphere without punctures. As a consequence, the properties of +tree-level string theory are found by studying the matter and ghost CFTs on the Riemann +sphere. Scattering amplitudes are computed through correlation functions of appropriate op- +erators on the sphere. This picture generalizes to higher-genus Riemann surfaces. Moreover, +since local properties of the CFT (e.g. the spectrum of operators) are determined by the +conformal algebra, they will be common to all surfaces. +Mathematically, a difference between Σ0 and Σ0,2 had to be expected since the sphere +has a positive curvature (and χ = −2) but the cylinder is flat (with χ = 0). Punctures +contribute negatively to the curvature (and thus positively to the Euler characteristics). +Remark 6.1 The coordinate z is always used as a coordinate on the complex plane, but the +corresponding metric may be different – compare (6.5) and (6.18). As explained previously, +this does not matter since the theory is insensitive to the conformal factor. +6.2 +Classical CFTs +In this section, we consider an action S[Ψ] which is conformally invariant. We first identify +and discuss the properties of the conformal algebra and group, before explaining how a CFT +is defined. +84 + +−−−−−−−−→ +Figure 6.3: Map from the cylinder with two caps (half-spheres) to the Riemann sphere Σ0. +6.2.1 +Witt conformal algebra +Since the Riemann sphere is identified with the complex plane, they share the same conformal +group and algebra. Consider the metric (6.5) +ds2 = dzd¯z, +(6.19) +then, any meromorphic change of coordinates +z −→ z′ = f(z), +¯z −→ ¯z′ = ¯f(¯z) +(6.20) +is a conformal transformation since the metric becomes +ds2 = dz′d¯z′ = +���� +df +dz +���� +2 +dzd¯z. +(6.21) +However, only holomorphic functions which are globally defined on ¯C are elements of +the group. At the algebra level, any holomorphic function f(z) regular in a domain D gives +a well-defined transformation in this domain D. Hence, the algebra is infinite-dimensional. +On the other hand, f(z) is only meromorphic on C generically: it cannot be exponentiated +to a group element. We first characterize the algebra and then obtain the conditions to +promote the local transformations to global ones. +Since the transformations are defined only locally, it is sufficient to consider an infinites- +imal transformation +δz = v(z), +δ¯z = ¯v(¯z), +(6.22) +where v(z) is a meromorphic vector field on the Riemann sphere. Indeed, the conformal +Killing equation (5.6) in D = 2 is equivalent to the Cauchy–Riemann equations: +¯∂v = 0, +∂¯v = 0. +(6.23) +The vector field admits a Laurent series +v(z) = +� +n∈Z +vnzn+1, +¯v(¯z) = +� +n∈Z +¯vn¯zn+1, +(6.24) +and the vn and ¯vn are to be interpreted as the parameters of the transformation. A basis of +vectors (generators) is: +ℓn = −zn+1∂z, +¯ℓn = −¯zn+1∂¯z, +n ∈ Z. +(6.25) +One can check that each set of generators satisfies the Witt algebra +[ℓm, ℓn] = (m − n)ℓm+n, +[¯ℓm, ¯ℓn] = (m − n)¯ℓm+n, +[ℓm, ¯ℓn] = 0. +(6.26) +85 + +Since there are two commuting copies of the Witt algebra, it is natural to extend the +ranges of the coordinates from C to C2 and to consider z and ¯z as independent variables. +In particular, this gives a natural action of the product algebra over C2. This procedure +will be further motivated when studying CFTs since the holomorphic and anti-holomorphic +parts will generally split, and it makes sense to study them separately. Ultimately, phys- +ical quantities can be extracted by imposing the condition ¯z = z∗ at the end (the star is +always reserved for the complex conjugation, the bar will generically denote an independent +variable). In that case, the two algebras are also related by complex conjugation. +Note that the variation of the metric (B.6) under a meromorphic change of coordinates +(6.22) becomes +δgz¯z = ∂v + ¯∂¯v, +δgzz = δg¯z¯z = 0. +(6.27) +6.2.2 +PSL(2, C) conformal group +The next step is to determine the globally defined vectors and to study the associated group. +First, the conditions for a vector v(z) to be well-defined at z = 0 are +lim +|z|→0 v(z) < ∞ +=⇒ +∀n < −1 : +vn = 0. +(6.28) +The behaviour at z = ∞ can be investigated thanks to the map z = 1/w +v(1/w) = dz +dw +� +n +vnw−n−1, +(6.29) +where the additional derivative arises because v is a vector. Then, the regularity conditions +at z = ∞ are +lim +|z|→∞ v(z) = lim +|w|→0 +dz +dw v(1/w) = − lim +|w|→0 +v(1/w) +w2 +< ∞ +=⇒ +∀n > 1 : +vn = 0. +(6.30) +As a result, the globally defined generators are +{ℓ−1, ℓ0, ℓ1} ∪ {¯ℓ−1, ¯ℓ0, ¯ℓ1} +(6.31) +where +ℓ−1 = −∂z, +ℓ0 = −z∂z, +ℓ1 = −z2∂z. +(6.32) +It is straightforward to check that they form two copies of the sl(2, C) algebra +[ℓ0, ℓ±1] = ∓ℓ±1, +[ℓ1, ℓ−1] = 2ℓ0. +(6.33) +The global conformal group is sometimes called Möbius group: +PSL(2, C) := SL(2, C)/Z2 ∼ SO(3, 1), +(6.34) +where the additional division by Z2 is clearer when studying an explicit representation. It +corresponds with ker P1 defined in (2.91): +K0 = PSL(2, C). +(6.35) +A matrix representation of SL(2, C) is +g = +�a +b +c +d +� +, +a, b, c, d ∈ C, +det g = ad − bc = 1, +(6.36) +86 + +which shows that this group has six real parameters +K0 := dim SL(2, C) = 6. +(6.37) +The associated transformation on the complex plane reads +fg(z) = az + b +cz + d. +(6.38) +The quotient by Z2 is required since changing the sign of all parameters does not change the +transformation. These transformations have received different names: Möbius, projective, +homographic, linear fractional transformations. . . +Holomorphic vector fields are then of the form +v(z) = β + 2αz + γz2, +¯v(¯z) = ¯β + 2¯α¯z + ¯γ¯z2, +(6.39) +where +a = 1 + α, +b = β, +c = −γ, +d = 1 − α. +(6.40) +The finite transformations associated to (5.12) are: +translation: +fg(z) = z + a, +a ∈ C, +(6.41a) +rotation: +fg(z) = ζ z, +|ζ| = 1, +(6.41b) +dilatation: +fg(z) = λ z, +λ ∈ R, +(6.41c) +SCT: +fg(z) = +z +cz + 1, +c ∈ C. +(6.41d) +Investigation leads to the following association between the generators and transformations: +• translation: ℓ−1 and ¯ℓ−1; +• dilatation (or radial translation): (ℓ0 + ¯ℓ0); +• rotation (or angular translation): i(ℓ0 − ¯ℓ0); +• special conformal transformation: ℓ1 and ¯ℓ1. +The inversion defined by +inversion: +I+(z) := I(z) := 1 +z +(6.42a) +is not an element of SL(2, C). However, the inversion with a minus sign +I−(z) := −I(z) = I(−z) = −1 +z +(6.42b) +is a SL(2, C) transformation. +A useful transformation is the circular permutation of (0, 1, ∞): +g∞,0,1(z) = +1 +1 − z . +(6.43) +87 + +6.2.3 +Definition of a CFT +A CFT is characterized by its set of (composite) fields (also called operators) O(z, ¯z) which +correspond to any local expression constructed from the fields Ψ appearing in the Lagrangian +and of their derivatives.5 For example, in a scalar field theory, the simplest operators are of +the form ∂mφn. +Among the operators, two particular categories are distinguished according to their trans- +formation laws: +• primary operator: +∀f meromorphic : +O(z, ¯z) = +�df +dz +�h �d ¯f +d¯z +�¯h +O′� +f(z), ¯f(¯z) +� +, +(6.44) +• quasi-primary (or SL(2, C) primary) operator: +∀f ∈ PSL(2, C) : +O(z, ¯z) = +�df +dz +�h �d ¯f +d¯z +�¯h +O′� +f(z), ¯f(¯z) +� +. +(6.45) +The parameters (h, ¯h) are the conformal weights of the operator O (both are independent +from each other), and combinations of them give the conformal dimension ∆ and spin s: +∆ := h + ¯h, +s := h − ¯h. +(6.46) +The conformal weights correspond to the charges of the operator under ℓ0 and ¯ℓ0. We will +use “(h, ¯h) (quasi-)primary” as a synonym of “(quasi-)primary field with conformal weight +(h, ¯h)”. +Remark 6.2 (Complex conformal weights) While we consider h, ¯h ∈ R, and more spe- +cifically h, ¯h ≥ 0 for a unitary theory (which is the case of string theory except for the re- +parametrization ghosts), theories with h, ¯h ∈ C make perfectly sense. One example is the +Liouville theory with complex central charge c ∈ C [200, 202] (central charges are defined +below, see (6.58)). +Primaries and quasi-primaries are hence operators which have nice transformations re- +spectively under the algebra and group. Obviously, a primary is also a quasi-primary. These +transformations are similar to those of a tensor with h holomorphic and ¯h anti-holomorphic +indices (Section 4.1). Another point of view is that the object +O(z, ¯z) dzhd¯z +¯h +(6.47) +is invariant under local / global conformal transformations. +The notation f ◦ O indicates the complete change of coordinates, including the tensor +transformation law and the possible corrections if the operator is not primary.6 For a primary +field, we have: +f ◦ O(z, ¯z) := f ′(z)h ¯f ′(¯z) +¯h O′� +f(z), ¯f(¯z) +� +. +(6.48) +We stress that it does not correspond to function composition. +Under an infinitesimal transformations +δz = v(z), +δ¯z = ¯v(¯z), +(6.49) +5Not all CFTs admit a Lagrangian description. But, since we are mostly interested in string theories +defined from Polyakov’s path integral, it is sufficient to study CFTs with a Lagrangian. +6In fact, one has f ◦ O := f∗O in the notations of Chapter 2. +88 + +a primary operator changes as +δO(z, ¯z) = (h ∂v + v ∂)O(z, ¯z) + (¯h ¯∂¯v + ¯v ¯∂)O(z, ¯z). +(6.50) +The transformation of a non-primary field contains additional terms, see for example (6.89). +Remark 6.3 (Higher-genus Riemann surfaces) According to Remark 5.1, all Riemann +surfaces Σg share the same conformal algebra since locally they are all subsets of R2. On +the other hand, one finds that no global transformations are defined for g > 1, and only the +subgroup U(1) × U(1) survives for the torus. +The most important operator in a CFT is the energy–momentum tensor Tµν, if it exists +as a local operator. According to Section 2.1, this tensor is conserved and traceless +∇νTµν = 0, +gµνTµν = 0. +(6.51) +The traceless equation in components reads +gµνTµν = 4 Tz¯z = Txx + Tyy = 0 +(6.52) +which implies that the off-diagonal component vanishes in complex coordinates +Tz¯z = 0. +(6.53) +Then, the conservation equation yields +∂zT¯z¯z = 0, +∂¯zTzz = 0, +(6.54) +such that the non-vanishing components Tzz and T¯z¯z are respectively holomorphic and anti- +holomorphic. This motivates the introduction of the notations: +T(z) := Tzz(z), +¯T(¯z) := T¯z¯z(¯z). +(6.55) +This is an example of the factorization between the holomorphic and anti-holomorphic sec- +tors. +Currents are local objects and thus one expects to be able to write an infinite number of +such currents associated to the Witt algebra. Applying the Noether procedure gives +Jv(z) := J ¯z +v (z) = −T(z)v(z), +¯Jv(¯z) := Jz +v (¯z) = − ¯T(¯z)¯v(¯z). +(6.56) +6.3 +Quantum CFTs +The previous section was purely classical. The quantum theory is first defined through the +path integral +Z = +� +dΨ e−S[Ψ]. +(6.57) +We will also develop an operator formalism. The latter is more general than the path integral +and allows to work without reference to path integrals and Lagrangians. This is particularly +fruitful as it extends the class of theories and parameter ranges (e.g. Remark 6.2) which can +be studied. +89 + +6.3.1 +Virasoro algebra +As discussed in Section 2.3.3, field measures in path integrals display a conformal anom- +aly, meaning that they cannot be defined without introducing a scale. This anomaly can +be traded for a gravitational anomaly by introducing counter-terms in the action [87, 95, +sec. 3.2, 96, 101, 122, 129]. As a consequence, the Witt algebra (6.26) is modified to its +central extension, the Virasoro algebra.7 +The generators in both sectors are denoted by +{Ln} and {¯Ln} and are called Virasoro operators (or modes). The algebra is given by: +[Lm, Ln] = (m − n)Lm+n + c +12 m(m − 1)(m + 1)δm+n, +(6.58a) +[¯Lm, ¯Ln] = (m − n)¯Lm+n + ¯c +12 m(m − 1)(m + 1)δm+n, +(6.58b) +[Lm, ¯Ln] = 0, +[c, Lm] = 0, +[¯c, ¯Lm] = 0, +(6.58c) +where c, ¯c ∈ C are the holomorphic and anti-holomorphic central charges. Consistency of +the theory on a curved space implies ¯c = c, but there is otherwise no constraint on the +plane [95, 246]. +The sl(2, C) subalgebra is not modified by the central extension. This means that states +are still classified by eigenvalues of (h, ¯h) of (L0, ¯L0). +Remark 6.4 In most models relevant for string theory, one finds that the central charges +are real, c, ¯c ∈ R. +Moreover, unitarity requires them to be positive c, ¯c > 0, and only +reparametrization ghosts do not satisfy this condition. On the other hand, it makes perfect +sense to discuss general CFTs for c, ¯c ∈ C (the Liouville theory is such an example [200, +202]). +6.3.2 +Correlation functions +A n-point correlation function is defined by +� n +� +i=1 +Oi(zi, ¯zi) +� += +� +dΨ e−S[Ψ] +n +� +i=1 +Oi(zi, ¯zi), +(6.59) +choosing a normalization such that ⟨1⟩ = 1. The path integral defines the time-ordered +product (on the cylinder) of the corresponding operators. +Invariance under global transformations leads to strong constraints on the correlation +functions. For quasi-primary fields, they transform under SL(2, C) as +� n +� +i=1 +Oi(zi, ¯zi) +� += +n +� +i=1 +�df +dz (zi) +�hi �df +d¯z (¯zi) +�¯hi +× +� n +� +i=1 +Oi +� +f(zi), ¯f(¯zi) +� +� +. +(6.60) +Considering an infinitesimal variation (6.50) yields a differential equation for the n-point +function +δ +� n +� +i=1 +Oi(zi, ¯zi) +� += +n +� +i=1 +� +hi∂iv(zi) + v(zi)∂i + c.c. +� +� n +� +i=1 +Oi(zi, ¯zi) +� += 0, +(6.61) +where ∂i := ∂zi and v is a vector (6.39) of sl(2, C). These equations are sufficient to determine +7That the central charge in the Virasoro algebra indicates a diffeomorphism anomaly can be understood +from the fact that +90 + +completely the forms of the 1-, 2- and 3-point functions of quasi-primaries: +⟨Oi(zi, ¯zi)⟩ = δhi,0δ¯hi,0, +(6.62a) +⟨Oi(zi, ¯zi)Oj(zj, ¯zj)⟩ = δhi,hjδ¯hi,¯hj +gij +z2hi +ij ¯z2¯hi +ij +, +(6.62b) +⟨Oi(zi, ¯zi)Oj(zj, ¯zj)Ok(zk, ¯zk)⟩ = +Cijk +zhi+hj−hk +ij +zhj+hk−hi +jk +zhi+hk−hj +ki +× +1 +¯z +¯hi+¯hj−¯hk +ij +¯z +¯hj+¯hk−¯hi +jk +¯z +¯hi+¯hk−¯hj +ki +, +(6.62c) +where we have defined +zij = zi − zj. +(6.63) +The coefficients Cijk are called structure constants and the matrix gij defines a metric +(Zamolodchikov metric) on the space of fields. The metric is often taken to be diagonal +gij = δij, which amounts to use an orthonormal eigenbasis of L0 and ¯L0. The vanishing of +the 1-point function of a non-primary quasi-primary holds only on the plane: for example +the value on the cylinder can be non-zero since the map is not globally defined – see in +particular (6.167). +Remark 6.5 (Logarithmic CFTs) Logarithmic CFTs display a set of unusual proper- +ties [84, 85, 90, 97, 130]. +In particular, the correlation functions are not of the form +displayed above. The most striking feature of those theories is that the L0 operator is non- +diagonalisable (but it can be set in the Jordan normal form). +Remark 6.6 (Fake identity) Usually, the only primary operator with h = ¯h = 0 is the +identity 1. While this is always true for unitary theories, there are non-unitary theories +(c ≤ 1 Liouville theory, SLE, loop models) where there is another field (called the indicator, +marking operator, or also fake identity) with h = ¯h = 0 [13, 46, 99, 112, 182, 200, 202]. +The main difference between both fields is that the identity is a degenerate field (it has a +null descendant), whereas the other operator with h = ¯h = 0 is not. Such theories will not +be considered in this book. Operators with h = ℏ = 0 can also be built by comining several +CFTs, and they play a very important role in string theory since they describe on-shell states. +Finally, the 4-point function is determined up to a function of a single variable x and its +complex conjugate: +� 4 +� +i=1 +Oi(zi, ¯zi) +� += f(x, ¯x) +� +i |w|, +(−1)F B(w)A(z) +|w| > |z|, +(6.72) +where F = 0 (F = 1) for bosonic (fermionic) operators. Radial ordering will often be kept +implicit. +The equal-time (anti-)commutator becomes an equal radius commutator defined by +point-splitting: +[A(z), B(w)]±,|z|=|w| = lim +δ→0 +� +A(z)B(w)||z|=|w|+δ ± B(w)A(z)||z|=|w|−δ +� +. +(6.73) +If A and B are two operators which can be written as the contour integrals of a(z) and b(z) +(corresponding to integral over closed curves on the cylinder) +A = +� +C0 +dz +2πi a(z), +B = +� +C0 +dz +2πi b(z), +(6.74) +then one finds the following commutators: +[A, B]± = +� +C0 +dw +2πi +� +Cw +dz +2πi a(z)b(w), +(6.75a) +[A, b(w)]± = +� +Cw +dz +2πi a(z)b(w). +(6.75b) +92 + +Figure 6.4: Graphical proof of (6.75). +The contours C0 and Cw are respectively centered around the points 0 and w. For a proof, +see Figure 6.4. Since these are contour integrals in the complex plane, the Cauchy–Riemann +formula (B.1) can be used to write the result as soon as one knows the poles of the above +expression (ultimately, this amounts to pick the sum of residues). In CFTs, the poles of such +expressions are given by operator product expansions (OPE), defined below (Section 6.4.2). +Given a conserved current jµ +∂µjµ = ∂jz + ¯∂j ¯z = 2(∂j¯z + ¯∂jz) = 0, +(6.76) +the associated conserved charge is defined by +Q = +1 +2πi +� +C0 +(jzdz − j¯zd¯z), +(6.77) +where C0 denotes the anti-clockwise contour around z = 0 (equivalently the interior of the +contour is located to the left). The difference of sign in the second term follows directly +from Stokes’ theorem (B.14g) (and can be understood as a conjugation of the contour). +The additional factor of 1/2π is consistent with the normalization of spatial integrals in two +dimensions. The current components are not necessarily holomorphic and anti-holomorphic +at this level, but in practice this will often be the case (and each component is independently +conserved), and one writes +j(z) := jz(z), +¯ȷ(¯z) := j¯z(¯z). +(6.78) +In this case, the charge also splits into a holomorphic and an anti-holomorphic (left- and +right-moving8) contributions +Q = QL + QR, +QL := +1 +2πi +� +C0 +j(z)dz, +QR := − 1 +2πi +� +C0 +¯ȷ(¯z)d¯z. +(6.79) +The infinitesimal variation of a field under the symmetry generated by Q reads +δϵO(z, ¯z) = −[ϵQ, O(z, ¯z)] = −ϵ +� +Cz +dw +2πi j(w)O(z, ¯z) + ϵ +� +Cz +d ¯w +2πi ¯ȷ( ¯w)O(z, ¯z). +(6.80) +The contour integrals are easily evaluated once the OPE between the current and the oper- +ator is known. This formula gives the infinitesimal variation under the transformation for +any field, not only for primaries. +8For charges, we use subscript L and R to distinguish both sectors to avoid introducing a new symbol for +the total charge. However, since Q = QL in the holomorphic sector, it is often not necessary to distinguish +between the two symbols when acting on an operator or a state (however, this is useful for writing mode +expansions). We do not write a bar on QR because the charges don’t depend on the position. +93 + +Computation – Equation (6.77) +In real coordinates, the charge is defined by integrating the time component of the +current jµ over space for fixed time (A.23): +Q = 1 +2π +� +dσ j0. +The first step is to rewrite this formula covariantly. Since the time is fixed on the slice, +dτ = 0 and one can write +Q = 1 +2π +� +(dσ j0 − dτ j1) = − 1 +2π +� +ϵµνjµ dxν. +The last formula is valid for any contour. Moreover, it can be evaluated for complex +coordinates: +Q = − 1 +2π +� +ϵz¯z +� +jz d¯z − j ¯z dz +� += − i +4π +� � +jz d¯z − j ¯z dz +� += − 1 +2πi +� � +jz dz − j¯z d¯z +� +. +One finds a contour integral because τ = cst circles of the cylinder are mapped to +|z| = cst contours. +6.4.2 +Operator product expansions +The operator product expansion (OPE) is a tool used frequently in CFT: it means that +when two local operators come close to each other, it is possible to replace their product by +a sum of local operators +Oi(zi, ¯zi)Oj(zj, ¯zj) = +� +k +ck +ij +zhi+hj−hk +ij +¯z +¯hi+¯hj−¯hk +ij +Ok(zj, ¯zj), +(6.81) +where the OPE coefficients ck +ij are some constants and the sum runs over all operators. +When Ok is primary, the coefficients ck +ij are related to the structure constants and the field +metric by +Cijk = gkℓcℓ +ij. +(6.82) +The radius of convergence for the OPE is given by the distance to the nearest operators in +the correlation function. The OPE defines an associative algebra (commutative for bosonic +operators), and the holomorphic sector forms a subalgebra (called the chiral algebra). +Example 6.1 – OPE with the identity +The OPE of a field φ(z) with the identity 1 is found by a direct series expansion +φ(z)1 = +� +n∈N +(z − w)n +n! +∂nφ(w). +(6.83) +Obviously there are no singular terms. +Starting from this point we consider only the holomorphic sector except when stated +otherwise. The formula for the OPE (6.81) can be rewritten as +A(z)B(w) := +N +� +n=−∞ +{AB}n(z) +(z − w)n +(6.84) +94 + +to simplify the manipulations. +N is an integer and there are singular terms if N > 0. +Generally, only the terms singular as w → z are necessary in the computations (for example, +to use the Cauchy–Riemann formula (B.1)): equality up to non-singular terms is denoted +by a tilde +A(z)B(w) ∼ +N +� +n=1 +{AB}n(z) +(z − w)n =: A(z)B(w). +(6.85) +The RHS of this expression defines the contraction of the operators A and B. +While, most of the time, only singular terms are kept +φi(zi)φj(zj) ∼ +� +k +θ(hi + hj − hk) +ck +ij +(z − w)hi+hj−hk φk(w) +(6.86) +(with θ(x) the Heaviside step function), it can happen that one keeps also non-singular terms +(the product of two OPE have singular terms coming from non-singular terms multiplying +singular terms). Explicit contractions of operators through the OPE is also denoted by a +bracket when there are other operators. +For a primary field φ(z), one finds the OPE with the energy–momentum tensor to be +T(z)φ(w) ∼ +h φ(w) +(z − w)2 + ∂φ(w) +z − w , +(6.87) +where h is the conformal weight of the field. This OPE together with (6.80) for j(z) = +−v(z)T(z) correctly reproduces (6.50). +Computation – Equation (6.50) +δφ(z) = +� +Cz +dw +2πi v(w)T(w)φ(z) ∼ +� +Cz +dw +2πi v(w) +� h φ(z) +(w − z)2 + ∂φ(z) +w − z +� += h ∂v(z) φ(z) + v(z)∂φ(z). +For a non-primary operator, the OPE becomes more complicated (as it is reflected by +the transformation property), but the conformal weight can still be identified at the term +in z−2. The most important example is the energy–momentum tensor: the central charge is +found as the coefficient of the z−4 term its OPE with itself: +T(z)T(w) ∼ +c/2 +(z − w)4 + +2T(w) +(z − w)2 + ∂T(w) +z − w . +(6.88) +The OPE indicates that the conformal weight of T is h = 2. +Using (6.80) for j(z) = +−v(z)T(z), one finds the infinitesimal variation +δT = 2 ∂v T + v ∂T + c +12 ∂3v, +(6.89) +The last term vanishes for global transformations: this translates the fact that T is only a +quasi-primary. The finite form of this transformation is +T ′(w) = +� dz +dw +�−2 � +T(z) − c +12 S(w, z) +� += +� dz +dw +�−2 +T(z) + c +12 S(z, w) +(6.90) +where S(w, z) is the Schwarzian derivative +S(w, z) = w(3) +w′ − 3 +2 +�w′′ +w′ +�2 +, +(6.91) +95 + +where the derivatives of w are with respect to z. This vanishes if the transformation is in +SL(2, C), and it transforms as +S(u, z) = S(w, z) + +�dw +dz +�2 +S(u, w) +(6.92) +under successive changes of coordinates. +Computation – Equation (6.89) +δT(z) = +� +Cz +dw +2πi v(w)T(w)T(z) ∼ +� +Cz +dw +2πi v(w) +� +c/2 +(z − w)4 + +2T(w) +(z − w)2 + ∂T(w) +z − w +� += +c +2 × 3! ∂3v(z) + 2∂v(z) T(z) + v(z)∂T(z). +6.4.3 +Hermitian and BPZ conjugation +In this section, we introduce two different notions of conjugations: one is adapted for amp- +litudes because it defines a unitary Euclidean time evolution, while the second is more +natural as an inner product of CFT states. Both can be interpreted as providing a map +from in-states to out-states on the cylinder. +Given an operator O, we need to define an operation O‡ – called Euclidean adjoint (or +simply adjoint) – which, after Wick rotation from Euclidean to Lorentzian signature, can +be interpreted as the Hermitian adjoint.9 This is necessary in order to define a Hermitian +inner-product and to impose reality conditions. +To motivate the definition, consider first the cylinder in Lorentzian signature. +Since +Hermitian conjugation does not affect the Lorentzian coordinates, the Euclidean time must +reverse its sign: +t† = −iτ † = t +=⇒ +τ † = −τ. +(6.93) +Hence, an appropriate definition of the Euclidean adjoint is an Hermitian conjugation to- +gether with time reversal.10 +Another point of view is that the time evolution operator +U(τ) := e−τH is not unitary when H is Hermitian H† = H: the solution is to define a new +Euclidean adjoint U(τ)‡ := U(−τ)† such that U(τ) is unitary for it. +Time reversal on the cylinder corresponds to inversion and complex conjugation on the +complex plane: +z +τ→−τ +−−−−→ e−τ+iσ = 1 +z∗ = I(¯z), +(6.94) +where I(z) = 1/z is the inversion (6.42).11 On the real surface12 ¯z = z∗, which leads to the +definition of the Euclidean adjoint as follows: +O(z, ¯z)‡ := +�¯I ◦ O(z, ¯z) +�†, +(6.95) +9In [193], it is denoted by a bar on top of the operator: we avoid this notation since the bar already +denotes the anti-holomorphic sector. In [262], it is indicated by a subscript hc. Otherwise, in most of the +literature, it has no specific symbol since one directly works with the modes. +10The Euclidean adjoint can be used to define an inner product: positive-definiteness of the latter is +called reflection positivity or OS-positive and is a central axiom of constructive QFT. +11We do not write “z†” because this notation is confusing as one should not complex conjugate the factor +of i in the exponential (Section 6.2.1). +12Remember that ¯z is not the complex conjugate of z but an independent variable. +96 + +where ¯I(z) := 1/¯z. If O is quasi-primary, we have: +O(z, ¯z)‡ = +� +1 +¯z2hz2¯h O +�1 +¯z , 1 +z +��† += +1 +z2h¯z2¯h O† +�1 +z , 1 +¯z +� +, +(6.96) +The last equality shows that Euclidean conjugation is equivalent to take the conjugate of all +factors of i but otherwise leaves z and ¯z unaffected. The Euclidean adjoint acts by complex +conjugation of any c-number and reverses the order of the operators (acting as a transpose): +(λ O1 · · · On)‡ = λ∗ O‡ +n · · · O‡ +1, +λ ∈ C, +(6.97) +without any sign. +A second operation, called the BPZ conjugation, is useful. +It can be defined in two +different ways: +O(z, ¯z)t := I± ◦ O(z, ¯z) = (∓1)h+¯h +z2h¯z2¯h O +� +±1 +z , ±1 +¯z +� +, +(6.98) +where I±(z) = ±1/z is the inversion (6.42). The minus and plus signs are respectively more +convenient when working with the open and closed strings.13 The BPZ conjugation does +not complex conjugate c-number nor changes the order of the operators:14 +(λ O1 · · · On)t = λ Ot +1 · · · Ot +n, +λ ∈ C. +(6.99) +The identity is invariant under both conjugation +1‡ = 1t = 1. +(6.100) +6.4.4 +Mode expansion +Any field of weight (h, ¯h) can be expanded in terms of modes Om,n +O(z, ¯z) = +� +m,n +Om,n +zm+h¯zn+¯h . +(6.101) +Note that the modes Om,n themselves are operators. The ranges of the two indices are such +that +m + h ∈ Z + ν, +n + ¯h ∈ Z + ¯ν, +ν, ¯ν = +� +0 +periodic, +1/2 +anti-periodic. +(6.102) +The values of ν and ¯ν depend on whether the fields satisfies periodic or anti-periodic bound- +ary conditions on the plane (for half-integer weights, the periodicity is reversed on the +cylinder): +O(e2πiz, ¯z) = e2πiνO(z, ¯z), +O(z, e2πi¯z) = e2πi¯νO(z, ¯z). +(6.103) +Depending on whether the weights are integers or half-integers, additional terminology is +introduced: +13The index t should not be confused with the matrix transpose: it is used in opposition with ‡ and † to +indicate that no complex conjugation is involved. +14However, the fields become anti-radially ordered after a BPZ conjugation since it sends z to 1/z. The +radial ordering can be restored by (anti-)commuting the fields, which can introduce additional signs [220]. +This problem does not arise when working in terms of the modes. +97 + +• If h ∈ Z + 1/2, then one can choose anti-periodic (Neveu–Schwarz or NS) or periodic +(Ramond or R) boundary conditions on the cylinder (reversed for the plane): +ν, ¯ν = +� +0 +NS +1/2 +R +(6.104) +The indices are half-integers (resp. integers) for the NS (R) sector. +• If h ∈ Z, periodic (or untwisted) boundary conditions are more natural, but anti- +periodic boundary conditions may also be considered: +ν, ¯ν = +� +0 +untwisted +1/2 +twisted +(6.105) +The modes of untwisted (resp. twisted) fields have integer (half-integers) indices. +The mode expansions have no branch cut (fractional power of z or ¯z) for periodic fields +(bosonic untwisted or fermionic twisted). We will see explicit examples of such operators +later in this book. +Under Euclidean conjugation (6.95), the modes are related by +(O‡)−m,−n = (Om,n)†. +(6.106) +In particular, if the operator is Hermitian (under the Euclidean adjoint), the reality condition +on the modes relates the negative modes with the conjugated positive modes +O‡ = O +=⇒ +(Om,n)† = O−m,−n. +(6.107) +When no confusion is possible (for Hermitian operators), we will write O† +m,n instead of +(Om,n)†. +For a holomorphic field φ(z), the above expansion becomes +φ(z) = +� +n∈Z+h+ν +φn +zn+h . +(6.108) +Conversely, the modes are recovered from the field through +φn = +� +C0 +dz +2πi zn+h−1φ(z), +(6.109) +where the integration is counter-clockwise around the origin. +If the field is Hermitian, then +φ‡ = φ +=⇒ +(φn)† = φ−n. +(6.110) +The operators φn have a conformal weight of −n (since the weight of z is −1). The BPZ +conjugate of the modes is +φt +n = (I± ◦ φ)n = (−1)h(±1)nφ−n. +(6.111) +98 + +Computation – Equation (6.111) +φt +n = (I± ◦ φ)n = +� +dz +2πi zn+h−1I± ◦ φ(z) += +� +dz +2πi zn+h−1 +� +∓ 1 +z2 +�h +φ +� +±1 +z +� += (∓1)h +� +dz +2πi zn−h−1φ +� +±1 +z +� += (∓1)h +� dw +2πi +� +± 1 +w +�n−h +w−1φ(w) += (∓1)h(±1)n−h +� dw +2πi w−n+h−1φ(w), +where we have set w = ±1/z such that +dz +z = ∓ dw +w2z = −dw +w , +(6.112) +and the minus sign disappears upon reversing the contour orientation. +The mode expansion of the energy–momentum tensor is +T(z) = +� +n∈Z +Ln +zn+2 , +Ln = +� +dz +2πi T(z)zn+1, +(6.113) +where one recognizes the Virasoro operators as the modes. In most situations, the Virasoro +operators are Hermitian +L† +n = L−n. +(6.114) +The OPE (6.88) and (6.87) together with (6.75a) help to reconstruct the Virasoro algebra +(6.58) and the commutation relations between the Lm and the modes φn of a weight h +primary: +[Lm, φn] = +� +m(h − 1) − n +� +φm+n. +(6.115) +This easily gives the commutation relation for the complete field: +[Lm, φ(z)] = zm� +z∂ + (n + 1)h +� +φ(z). +(6.116) +We will often use (6.58) and (6.115) for m = 0: +[L0, L−n] = nL−n, +[L0, φ−n] = nφ−n. +(6.117) +This means that both φn and Ln act as raising operators for L0 if n < 0, and as lowering +operators if n > 0 (remember that L0 is the Hamiltonian in the holomorphic sector). When +both the holomorphic and anti-holomorphic sectors enter, it is convenient to introduce the +combinations +L± +n = Ln ± ¯Ln, +(6.118) +such that L+ +0 is the Hamiltonian. +Finally, every holomorphic current j(z) has a conformal weight h = 1 and can be expan- +ded as +j(z) = +� +n +jn +zn+1 . +(6.119) +By definition, the zero-mode is equal to the holomorphic charge +QL = j0. +(6.120) +99 + +6.4.5 +Hilbert space +The Hilbert space of the CFT is denoted by H. The SL(2, C) (or conformal) vacuum15 |0⟩ +is defined by the state which is invariant under the global conformal transformations: +L0 |0⟩ = 0, +L±1 |0⟩ = 0. +(6.121) +Expectation value of an operator O in the SL(2, C) vacuum is denoted as: +⟨O⟩ :=⟨0| O |0⟩ . +(6.122) +If the fields are expressed in terms of creation and annihilation operators (which happens +e.g. for free scalars, free fermions and ghosts), then the Hilbert space has the structure of a +Fock space. +State–operator correspondence +The state–operator correspondence identifies every state |O⟩ of the CFT Hilbert space with +an operator O(z, ¯z) through +|O⟩ = lim +z,¯z→0 O(z, ¯z) |0⟩ = O(0, 0) |0⟩ . +(6.123) +Such a state can be interpreted as an “in” state since it is located at τ → −∞ on the +cylinder. Focusing now on a holomorphic field φ(z), the state is defined as +|φ⟩ = lim +z→0 φ(z) |0⟩ = φ(0) |0⟩ . +(6.124) +For this to make sense, the modes which diverge as z → 0 must annihilate the vacuum. In +particular, for a weight h field φ(z), one finds: +∀n ≥ −h + 1 : +φn |0⟩ = 0. +(6.125) +Thus, the φn for n ≥ −h + 1 are annihilation operators for the vacuum |0⟩, and conversely +the states φn with n < −h + 1 are creation operators. As a consequence, the state |φ⟩ is +found by applying the mode n = −h to the vacuum: +|φ⟩ = φ−h |0⟩ = +� +dz +2πi +φ(z) +z +|0⟩ . +(6.126) +Since L−1 is the generator of translations on the plane, one finds +φ(z) |0⟩ = ezL−1φ(0)e−zL−1 |0⟩ = ezL−1 |φ⟩ . +(6.127) +The vacuum |0⟩ is the state associated to the identity 1. Translating the conditions (6.125) +to the energy–momentum tensor gives +∀n ≥ −1 : +Ln |0⟩ = 0. +(6.128) +This is consistent with the definition (6.121) since it includes the sl(2, C) subalgebra. +If h < 0, some of the modes with n > 0 do not annihilate the vacuum: (6.117) implies that +some states have an energy lower than the one of |0⟩. The state |Ω⟩ (possibly degenerate) +with the lowest energy is called the energy vacuum +∀ |φ⟩ ∈ H : +⟨Ω| L0 |Ω⟩ ≤⟨φ| L0 |φ⟩ . +(6.129) +15There are different notions of “vacuum”, see (6.129). However, the SL(2, C) vacuum is unique. Indeed, +it is mapped to the unique identity operator under the state–operator correspondence (however, there can +be other states of weight 0, see Remark 6.6). +100 + +It is obtained by acting repetitively with the modes φn>0. +This vacuum defines a new +partition of the non-zero-modes operators into annihilation and creation operators. If there +are zero-modes, i.e. n = 0 modes, then the vacuum is degenerate since they commute +with the Hamiltonian, [L0, φ0] = 0 according to (6.117). The partition of the zero-modes +into creation and annihilation operators depends on the specific state chosen among the +degenerate vacua. +The energy aΩ of |Ω⟩, which is also its L0 eigenvalue +L0 |Ω⟩ := aΩ |Ω⟩ , +(6.130) +is called zero-point energy. Bosonic operators with negative h are dangerous because they +lead to an infinite negative energy together with an infinite degeneracy (from the zero-mode). +The conjugate vacuum is defined by BPZ or Hermitian conjugation +⟨0| = |0⟩‡ = |0⟩t +(6.131) +since both leave the identity invariant. It is also annihilated by the sl(2, C) subalgebra: +⟨0| L0 = 0, +⟨0| L±1 = 0. +(6.132) +Since there are two kinds of conjugation, two different conjugated states can be defined. They +are also called “out” states since they are located at τ → ∞ on the cylinder (Figure 6.2). +Euclidean and BPZ conjugations and inner products +The Euclidean adjoint ⟨O‡| of the state |O⟩ is defined as +⟨O‡| = +lim +w, ¯ +w→0⟨0| O(w, ¯w)‡ = +lim +w, ¯ +w→0 +1 +w2h ¯w2¯h ⟨0| O +� 1 +¯w, 1 +w +�† +(6.133a) += +lim +z,¯z→∞ z2h¯z2¯h⟨0| O†(z, ¯z) +(6.133b) +=⟨0| I ◦ O†(0, 0), +(6.133c) +where the two coordinate systems are related by w = 1/¯z. From this formula, the definition +of the adjoint of a holomorphic operator φ follows +⟨φ‡| = lim +¯ +w→0⟨0| φ(w)‡ = lim +¯ +w→0 +1 +w2h ⟨0| φ† +� 1 +w +� +(6.134a) += lim +z→∞ z2h⟨0| φ†(z) +(6.134b) +=⟨0| I ◦ φ†(0). +(6.134c) +Then, expanding the field in terms of the modes gives +⟨φ‡| =⟨0| (φ†)h. +(6.135) +The BPZ conjugated state is +⟨φ| := lim +w→0⟨0| φ(w)t +(6.136a) += (±1)h lim +z→∞ z2h⟨0| φ(z) +(6.136b) +=⟨0| I± ◦ φ(0). +(6.136c) +In terms of the modes, one has +⟨φ| = (±1)h⟨0| φh. +(6.137) +101 + +If φ is Hermitian, then the relation between both conjugated states corresponds to a reality +condition: +⟨φ‡| = (±1)h⟨φ| . +(6.138) +Taking the BPZ conjugation of the conditions (6.125) tells which modes must annihilate +the conjugate vacuum: +∀n ≤ h − 1 : +⟨0| φn = 0, +(6.139) +and one finds more particularly for the Virasoro operators +∀n ≤ 1 : +⟨0| Ln = 0. +(6.140) +This can also be derived directly from (6.136) by requiring that applying an operator on the +conjugate vacuum ⟨0| is well-defined. +All conditions taken together mean that the expectation value of the energy–momentum +tensor in the conformal vacuum vanishes: +⟨0| T(z) |0⟩ = 0. +(6.141) +In particular, this means that the energy vacuum |Ω⟩, if different from |0⟩, has a negative +energy. +The Hermitian16 and BPZ inner products are respectively defined by: +⟨φ‡ +i|φj⟩ =⟨0| ¯I ◦ φj(0)φi(0) |0⟩ = lim +z→∞ +w→0 +z2hi⟨0| φ† +i(z)φj(w) |0⟩ , +(6.142a) +⟨φi|φj⟩ =⟨0| I ◦ φj(0)φi(0) |0⟩ = (±1)hi lim +z→∞ +w→0 +z2hi⟨0| φi(z)φj(w) |0⟩ . +(6.142b) +These products can be recast as 2-point correlation functions (6.62b) on the sphere: +⟨φi|φj⟩ = ⟨I ◦ φi(0)φj(0)⟩, +⟨φ‡ +i|φj⟩ = ⟨I ◦ φ† +i(0)φj(0)⟩. +(6.143) +From the state–operator correspondence, the action of one operator on the in-state can be +reinterpreted as the matrix element of this operator using the two external states, or also as +a 3-point function: +⟨φi| φj(z) |φk⟩ = (±1)hi lim +w→∞ w2hi⟨φi(w)φj(z)φk(0)⟩. +(6.144) +Given a basis of states {φi} (i can run over both discrete and continuous indices), the +conjugate or dual states {φc +i} are defined by: +⟨φc +i|φj⟩ = δij +(6.145) +(the delta function is discrete and/or continuous according to the indices). +Verma modules +If φ(z) is a weight h primary, then the associated state |φ⟩ satisfies: +L0 |φ⟩ = h |φ⟩ , +∀n ≥ 1 : +Ln |φ⟩ = 0. +(6.146) +Such a state is also called a highest-weight state. The descendant states are defined by all +possible states of the form +|φ{ni}⟩ := +� +i +L−ni |φ⟩ , +(6.147) +where the same L−ni can appear multiple times and ni > 0. The set of states φ{ni} is called +a Verma module V (h, c). One finds that the L0 eigenvalues of this state is +L0 = h + +� +i +ni. +(6.148) +16Depending on the normalization, it can also be anti-Hermitian. +102 + +Normal ordering +The normal ordering of an operator with respect to a vacuum corresponds to placing all cre- +ation (resp. annihilation) operators of this vacuum on the left (resp. right). From this defin- +ition, the expectation value of a normal ordered operator in the vacuum vanishes identically. +The main reason for normal ordering is to remove singularities in expectation values. +Given an operator φ(z), we define two normal orderings: +• The conformal normal order (CNO) :O: is defined with respect to the conformal va- +cuum (6.121): +⟨0| :O: |0⟩ = 0. +(6.149) +• The energy normal order (ENO) +⋆ +⋆ O +⋆ +⋆ is defined with respect to the energy vacuum +(6.129): +⟨Ω| +⋆ +⋆O +⋆ +⋆ |Ω⟩ = 0. +(6.150) +We first discuss the conformal normal ordering before explaining how to relate it to the +energy normal ordering. +Given two operators A and B, the simplest normal ordering amounts to subtract the +expectation value: +:A(z)B(w): +?= A(z)B(w) − ⟨A(z)B(w)⟩. +(6.151) +This is equivalent to defining the products of two operators at coincident points via point- +splitting: +:A(z)B(z): +?= lim +w→z +� +A(z)B(w) − ⟨A(z)B(w)⟩ +� +. +(6.152) +While this works well for free fields, this does not generalize for composite or interacting +fields. +The reason is that this procedure removes only the highest singularity in the product: it +does not work if the OPE has more than one singular term. An appropriate definition is +:A(z)B(w): := A(z)B(w) − A(z)B(w) = +� +n∈N +(z − w)n{AB}−n(z), +(6.153) +where the contraction between A and B is defined in (6.85), and the second equality comes +from (6.84). +Then, the product evaluated at coincident points is found by taking the limit (in this +case the argument is often indicated only at the end of the product) +:AB(z): := :A(z)B(z): := lim +w→z :A(z)B(w): = {AB}0(z). +(6.154) +Indeed, since all powers of (z − w) are positive in the RHS of (6.153), all terms but the first +one disappear. The form of (6.154) shows that the normal order can also be computed with +the contour integral +:AB(z): = +� +Cz +dw +2πi +A(z)B(w) +z − w +. +(6.155) +It is common to remove the colons of normal ordering when there is no ambiguity and, in +particular, to write: +AB(z) := :AB(z):. +(6.156) +In terms of modes, one has +:AB(z): = +� +m +:AB:m +zm+hA+hB , +(6.157a) +:AB:m = +� +n≤−hA +AnBm−n + +� +n>−hA +Bm−nAn. +(6.157b) +103 + +This expression makes explicit that normal ordering is non-commutative and non-associative: +:AB(z): ̸= :BA(z):, +:A(BC)(z): ̸= :(AB)C(z):. +(6.158) +The product of normal ordered operators can then be computed using Wick theorem. In +fact, one is more interested in the contraction of two such operators in order to recover the +OPE between these operators: the product is then derived with (6.153). +If Ai (i = 1, 2, 3) are free fields, one has +A1(z) :A2A3(w): = :A1(z)A2A3(w): + A1(z) :A2 A3(w):, +A1(z) :A2 A3(w): = A1(z)A2(w) :A3(w): + A1(z)A3(w) :A2(w):. +(6.159) +If the fields are not free, then the contraction cannot be extracted from the normal ordering. +Similarly if there are more fields, then one needs to perform all the possible contractions. +Given two free fields A and B, one has the following identities: +A(z) :B(w)n: = n A(z)B(w) :B(w)n−1:, +(6.160a) +A(z) :eB(w): = A(z)B(w) :eB(w):, +(6.160b) +:eA(z): :eB(w): = exp +� +A(z)B(w) +� +:eA(z)eB(w):. +(6.160c) +The last relation generalizes for a set of n fields Ai: +n +� +i=1 +:eAi: = : exp +� n +� +i=1 +Ai +� +: exp +� +i0 +Bm−nAn. +(6.162b) +To simplify the definition we assume that A0 is a creation operator and it is thus included +in the first sum (this must be adapted in function of which vacuum state is chosen if the +latter is degenerate). +The relation between the normal ordered modes is +:AB:m = +⋆ +⋆AB +⋆ +⋆m + +hA−1 +� +n=0 +[Bm+n, A−n]. +(6.163) +Computation – Equation (6.163) +:AB:m = +� +n≤−hA +AnBm−n + +� +n>−hA +Bm−nAn += +� +n≥hA +A−nBm+n + +� +n>0 +Bm−nAn + +hA−1 +� +n=0 +Bm+nA−n += +� +n≥0 +A−nBm+n + +� +n>0 +Bm−nAn + +hA−1 +� +n=0 +[Bm+n, A−n] += +⋆ +⋆AB +⋆ +⋆m + +hA−1 +� +n=0 +[Bm+n, A−n]. +The choice of the normal ordering for the operators is related to the ordering ambiguity +when quantizing the system: when the product of two non-commuting modes appears in the +classical composite field, the corresponding quantum operator is ambiguous (generally up to +a constant). In practice, one starts with the conformal ordering since it is invariant under +conformal transformations and because one can compute with contour integrals. Then, the +expression can be translated in the energy ordering using (6.163). But, knowing how the +conformal and energy vacua are related, it is often simpler to find the difference between +the two orderings by applying the operator on the vacua. +6.4.6 +CFT on the cylinder +According to (6.44), the relation between the field on the cylinder and on the plane is +φ(z) = +� L +2π +�h +z−hφcyl(w) +(6.164) +(quantities without indices are on the plane by definition). The mode expansion on the +cylinder is +φcyl = +�2π +L +�h � +n∈Z +φne− 2π +L w = +�2π +L +�h � +n∈Z +φn +zn . +(6.165) +105 + +Using the finite transformation (6.90) for the energy–momentum tensor T, one finds the +relation +Tcyl(w) = +�2π +L +�2 � +T(z)z2 − c +24 +� +. +(6.166) +For the L0 mode, one finds +(L0)cyl = L0 − c +24, +(6.167) +and thus the Hamiltonian is +H = (L0)cyl + (¯L0)cyl = L0 + ¯L0 − c + ¯c +24 . +(6.168) +6.5 +Suggested readings +• The most complete reference on CFTs is [54] but it lacks some recent developments. +Two excellent complementary books are [25, 206]. +String theory books generally dedicate a fair amount of pages to CFTs: particularly +good summaries can be found in [24, 128, 193, 194]. +Finally, a modern and fully algebraic approach can be found in [200, 201]. Other good +reviews are [196, 259]. +• There are various other books [104, 120, 126, 171] and reviews [32, 89, 91, 204, 243]. +• The maps from the sphere and the cylinder to the complex plane are discussed in [193, +sec. 2.6, 6.1]. +• Normal ordering is discussed in details in [54, chap. 6] (see also [24, sec. 4.2, 193, +sec. 2.2]). +• Euclidean conjugation is discussed in [54, sec. 6.1.1, 193, p. 202–3]. For a comparison +of Euclidean and BPZ conjugations, see [262, sec. 2.2, 220, p. 11]. +• Normal ordering and difference between the different definitions are described in [193, +chap. 2, 54, sec. 6.5]. +106 + +Chapter 7 +CFT systems +Abstract +This chapter summarizes the properties of some CFT systems. We focus on +the free scalar field and on the first-order bc system (which generalizes the reparametriz- +ation ghosts). For the different systems, we first provide an analysis on a general curved +background before focusing on the complex plane. This is sufficient to describe the local +properties on all Riemann surfaces g ≥ 0. +7.1 +Free scalar +7.1.1 +Covariant action +The Euclidean action of a free scalar X on a curved background gµν is +S = +ϵ +4πℓ2 +� +d2x√g gµν∂µX∂νX, +(7.1) +where ℓ is a length scale1 and +ϵ := +� ++1 +spacelike +−1 +timelike , +√ϵ := +� ++1 +spacelike +i +timelike +(7.2) +denotes the signature of the kinetic term. The field is periodic along σ +X(τ, σ) ∼ X(τ, σ + 2π). +(7.3) +The energy–momentum tensor reads +Tµν = − ϵ +ℓ2 +� +∂µX∂νX − 1 +2 gµν(∂X)2 +� +, +(7.4) +and it is traceless +T µ +µ = 0. +(7.5) +The equation of motion is +∆X = 0, +(7.6) +where ∆ is the Laplacian (A.28). +1To be identified with the string scale, such that α′ = ℓ2. +107 + +The simplest method for finding the propagator in flat space is by using the identity +(assuming that there is no boundary term) +0 = +� +dX +δ +δX(σ) +� +e−S[X]X(σ′) +� +, +(7.7) +which yields a differential equation for the propagator: +⟨∂2X(σ)X(σ′)⟩ = −2πϵℓ2 δ(2)(σ − σ′). +(7.8) +This is easily integrated to +⟨X(σ)X(σ′)⟩ = −ϵℓ2 +2 +ln |σ − σ′|2. +(7.9) +Computation – Equation (7.9) +By translation and rotation invariance, one has +⟨X(σ)X(σ′)⟩ = G(r), +r = |σ − σ′|. +(7.10) +In polar coordinates, the Laplacian reads +∆G(r) = 1 +r ∂r(rG′(r)). +(7.11) +Integrating the differential equation (7.8) over d2σ = rdrdθ yields +−2πϵℓ2 = 2π +� r +0 +dr′ r′ × 1 +r′ ∂r′(r′G′(r′)) = 2πrG′(r). +(7.12) +The solution is +G′(r) = −ϵℓ2 ln r +(7.13) +and the form (7.9) follows by writing +ln r = 1 +2 ln r2 = 1 +2 ln |σ − σ′|2. +(7.14) +The action (7.1) is obviously invariant under constant translations of X: +X −→ X + a, +a ∈ R. +(7.15) +The associated U(1) current2 is conserved and reads +Jµ := 2πiϵ +∂L +∂(∂µX) = i +ℓ2 gµν∂νX, +∇µJµ = 0. +(7.16) +On flat space, the charge follows from (A.23): +p = 1 +2π +� +dσ J0 = +i +2πℓ2 +� +dσ ∂0X. +(7.17) +This charge is called momentum because it corresponds to the spacetime momentum in +string theory. +2The group is R but the algebra is u(1) (since locally there is no difference between the real line and the +circle). +108 + +Moreover, there is a another topological current +�Jµ := −i ϵµνJν = 1 +ℓ2 ϵµν∂νX, +(7.18) +which is identically conserved: +∇µ �Jµ ∝ ϵµν[∇µ, ∇ν]X = 0 +(7.19) +since [∇µ, ∇ν] = 0 when acting on a scalar field. Note that ˜Jµ is the Hodge dual of Jµ. The +conserved charge is called the winding number and reads on flat space: +w = 1 +2π +� +dσ �J0 = +1 +2πℓ2 +� 2π +0 +dσ ∂1X = +1 +2πℓ2 +� +X(τ, 2π) − X(τ, 0) +� +. +(7.20) +Remark 7.1 (Normalization of the current) The definition of the current (7.16) may +look confusing. The factor of i is due to the Euclidean signature, see (A.25a), and the factor +of 2π comes from the normalization of the spatial integral. We have inserted ϵ in order to +interpret the conserved charge p as a component of the momentum contravariant vector in +string theory. +To make contact with string theory, consider D scalar fields Xa(xµ). Then, the current +becomes +Jµ +a = +i +2πℓ2 ηab∂µXb, +(7.21) +where the position of the indices is in agreement with the standard form of Noether’s formula +(A.25a) (a current has indices in opposite locations as the parameters and fields). Since we +have η00 = −1 = ϵX0, we find that J0µ = ϵX0Jµ +0 has no epsilon after replacing the expression +(7.17) of Jµ +0 . +The transformation Xa → Xa +ca is a global translation in target spacetime: the charge +pa is identified with the spacetime momentum. +The factor of i indicates that pa is the +Euclidean contravariant momentum vector by comparison with (A.7). +The convention of this section is to always work with quantities which will become con- +travariant vector to avoid ambiguity. +7.1.2 +Action on the complex plane +In complex coordinates, the action on flat space reads +S = +ϵ +2πℓ2 +� +dzd¯z ∂zX∂¯zX, +(7.22) +giving the equation of motion: +∂z∂¯zX = 0. +(7.23) +This indicates that ∂zX and ∂¯zX are respectively holomorphic and anti-holomorphic such +that +X(z, ¯z) = XL(z) + XR(¯z), +(7.24) +and we will remove the subscripts when there is no ambiguity (for example, when the position +dependence is written): +X(z) := XL(z), +X(¯z) := XR(¯z). +(7.25) +It looks like XL(z) and XR(¯z) are unrelated, but this is not the case because of the zero- +mode, as we will see below. +109 + +The U(1) current is written as +J := Jz = i +ℓ2 ∂zX, +¯J := J¯z = i +ℓ2 ∂¯zX, +(7.26) +where we used the relations Jz = J ¯z/2 and J¯z = Jz/2. The equation of motion implies that +the current J is holomorphic, and ¯J is anti-holomorphic: +¯∂J = 0, +∂ ¯J = 0. +(7.27) +The momentum splits into left- and right-moving parts: +p = pL + pR, +pL = +1 +2πi +� +dz J, +pR = − 1 +2πi +� +d¯z ¯J. +(7.28) +The components of the topological current (7.18) are related to the ones of the U(1) +current: +�Jz = i +ℓ2 ∂zX = J, +�J¯z = − i +ℓ2 ∂¯zX = − ¯J. +(7.29) +As a consequence, the winding number is +w = pL − pR. +(7.30) +Note that we have the relations +pL = p + w +2 +, +pR = p − w +2 +, +(7.31a) +p2 + w2 = p2 +L + p2 +R, +2pw = p2 +L − p2 +R. +(7.31b) +The energy–momentum tensor is +T := Tzz = − ϵ +ℓ2 ∂zX∂zX, +¯T := T¯z¯z = − ϵ +ℓ2 ∂¯zX∂¯zX, +Tz¯z = 0. +(7.32) +Since the ∂zX (∂¯zX) is (anti-)holomorphic, so is T(z) ( ¯T(¯z)). Since the energy–momentum +tensor, the current and the field itself (up to zero-modes) split in holomorphic and anti- +holomorphic components in a symmetric way, it is sufficient to focus on one of the sectors, +say the holomorphic one. +The other primary operators of the theory are given by the vertex operators Vk(z):3 +Vk(z, ¯z) := :eiϵkX(z,¯z):. +(7.33) +Remark 7.2 In fact, it is possible to introduce more general vertex operators +VkL,kR(z, ¯z) := :e2iϵ +� +kLX(z)+kRX(¯z) +� +:, +(7.34) +but we will not consider them in this book. +Remark 7.3 (Plane and cylinder coordinates) The action in w-coordinate (cylinder) +takes the same form as a result of the conformal invariance of the scalar field, which in +practice results from the cancellation between the determinant and inverse metric. As a +consequence, every quantity derived from the classical action (equation of motion, energy– +momentum tensor. . . ) will have the same form in both coordinate systems: we will focus +on the z-coordinate, writing the w-coordinate expression when it is insightful to compare. +This is not anymore the case at the quantum level: anomalies may translate into differences +between quantities: to differentiate between the plane and cylinder quantities an index “cyl” +will be added when necessary (by convention, all quantities without qualification are on the +plane). +3The ϵ in the exponential is consistent with interpreting X and k as a contravariant vector. +110 + +7.1.3 +OPE +The OPE between X and itself is directly found from the propagator: +X(z)X(w) ∼ −ϵℓ2 +2 +ln(z − w). +(7.35) +By successive derivations, one finds the OPE between X and ∂X +∂X(z)X(w) ∼ −ϵℓ2 +2 +1 +z − w, +(7.36) +and between ∂X with itself +∂X(z)∂X(w) ∼ −ϵℓ2 +2 +1 +(z − w)2 . +(7.37) +The invariance under the permutation of z and w reflects the fact that X is bosonic and +that both operators in (7.37) are identical. +The OPE between ∂X and T allows to verify that the field ∂X is primary with h = 1: +T(z)∂X(w) ∼ ∂X(w) +(z − w)2 + ∂ +� +∂X(w) +� +z − w +. +(7.38) +The OPE of T with itself gives +T(z)T(w) ∼ 1 +2 +1 +(z − w)4 + +2T(w) +(z − w)2 + ∂T(w) +z − w +(7.39) +which shows that the central charge is +c = 1. +(7.40) +One finds that the operator ∂nX has conformal weight +h = n +(7.41) +since the OPE with T is +T(z)∂nX(w) ∼ · · · + n ∂nX(w) +(z − w)2 + ∂(∂nX(w)) +z − w +(7.42) +where the dots indicate higher negative powers of (z − w). These states are not primary for +n ≥ 2. Explicitly, for n = 2, one finds +T(z)∂2X(w) ∼ 2 ∂X(w) +(z − w)3 + +2 ∂2X +(z − w)2 + ∂(∂2X(w)) +z − w +. +(7.43) +The OPE of a vertex operator with the current J is +J(z)Vk(w, ¯w) ∼ ℓ2k +2 +Vk(w, ¯w) +z − w +. +(7.44) +This shows that the vertex operators Vk are eigenstates of the U(1) holomorphic current +with the eigenvalue given by the momentum (with a normalization of ℓ2). Then, the OPE +with T: +T(z)Vk(w, ¯w) ∼ hk Vk(w, ¯w) +(z − w)2 ++ ∂Vk(w, ¯w) +z − w +(7.45) +111 + +together with its anti-holomorphic counterpart show that the Vk are primary operators with +weight +(hk, ¯hk) = +�ϵℓ2k2 +4 +, ϵℓ2k2 +4 +� +, +∆k = ϵℓ2k2 +2 +, +sk = 0. +(7.46) +Note that classically hk = 0 since ℓ ∼ ℏ [246, p. 81]. The weight is invariant under k → −k. +Finally, the OPE between two vertex operators is +Vk(z, ¯z)Vk′(w, , ¯w) ∼ +Vk+k′(w, ¯w) +(z − w)−ϵkk′ℓ2/2 , +(7.47) +where only the leading term (non-necessarily singular) is displayed. In particular, correlation +functions should be computed for ϵkk′ < 0 in order to avoid exponential growth. +Computation – Equation (7.38) +T(z)∂X(w) = − ϵ +ℓ2 :∂X(z)∂X(z): ∂X(w) ∼ −2ϵ +ℓ2 :∂X(z)∂X(z): ∂X(w) ∼ +∂X(z) +(z − w)2 . +The result (7.38) follows by Taylor expanding the numerator. +Computation – Equation (7.39) +T(z)∂X(w) = 1 +ℓ4 :∂X(z)∂X(z): :∂X(w)∂X(w): +∼ 1 +ℓ4 +� +:∂X(z)∂X(z): :∂X(w)∂X(w): + :∂X(z)∂X(z): :∂X(w)∂X(w): ++ :∂X(z)∂X(z): :∂X(w)∂X(w): + perms +� +∼ 2 × 1 +4 +1 +(z − w)4 − 4 × 1 +2ℓ2 +1 +(z − w)2 :∂X(z)∂X(w): +∼ 1 +2 +1 +(z − w)4 − 2 +ℓ2 +1 +(z − w)2 +� +:∂X(w)∂X(w): + (z − w) :∂2X(w)∂X(w): +� +. +Computation – Equation (7.42) +T(z)∂nX(w) ∼ ∂n−1 +w +∂X(z) +(z − w)2 +∼ n! +∂X(z) +(z − w)n+1 +∼ +n! +(z − w)n+1 +� +· · · + +1 +(n − 1)! (z − w)n−1∂n−1(∂X(w)) ++ 1 +n! (z − w)n∂n(∂X(w)) +� +. +112 + +Computation – Equation (7.44) +Using (6.160b), one has: +∂X(z)Vk(w, ¯w) ∼ iϵk ∂X(z)X(w) Vk(w, ¯w) ∼ iϵk +� +−ϵℓ2 +2 +1 +z − w +� +Vk(w, ¯w). +Computation – Equation (7.45) +T(z)Vk(w, ¯w) ∼ − ϵ +ℓ2 :∂X(z)∂X(z): :eiϵkX(w, ¯ +w): +∼ iϵk +2 +1 +z − w ∂X(z) :eiϵkX(w, ¯ +w): − ϵ +ℓ2 ∂X(z) :∂X(z)eiϵkX(w, ¯ +w): +∼ iϵk +2 +1 +z − w +� +:∂X(z) eiϵkX(w, ¯ +w): + ∂X(z) :eiϵkX(w, ¯ +w): +� ++ iϵk +2 +:∂X(z)eiϵkX(w, ¯ +w): +z − w +∼ ϵk2ℓ2 +4 +Vk(w, ¯w) +(z − w)2 + iϵk :∂X(w)eiϵkX(w, ¯ +w): +z − w +. +In the first line, we consider a single contraction (hence, there is no factor of 2): the +reason is that considering the contractions symmetrically and not successively counts +twice the first term of the last line. Indeed, there is only one way to generate this term. +It is also possible to achieve the same result by expanding the exponential. +Computation – Equation (7.47) +Using (6.160c) and keeping only the leading term, one has: +Vk(z, ¯z)Vk′(w, ¯w) ∼ exp +� +− kk′ X(z, ¯z)X(w, ¯w) +� +:eiϵkX(z,¯z)eiϵk′X(w, ¯ +w): +∼ (z − w)ϵkk′ℓ2/2 Vk+k′(w, ¯w). +7.1.4 +Mode expansions +Since ∂X is holomorphic and of weight h = 1, it can be expanded as:4 +∂X = −i +� +ℓ2 +2 +� +n∈Z +αn z−n−1, +¯∂X = −i +� +ℓ2 +2 +� +n∈Z +¯αn ¯z−n−1, +(7.48) +where an individual mode can be extracted with a contour integral: +αn = i +� +dz +2πi zn−1∂X(z), +¯αn = i +� +dz +2πi zn−1 ¯∂X(z). +(7.49) +4The Fourier expansion is taken to be identical for ϵ = ±1 fields since ∂X is contravariant in target +space. The difference between the two cases will appear in the commutators. +113 + +Integrating this formula gives: +X(z) = xL +2 − i +� +ℓ2 +2 α0 ln z + i +� +ℓ2 +2 +� +n̸=0 +αn +n z−n, +X(¯z) = xR +2 − i +� +ℓ2 +2 ¯α0 ln ¯z + i +� +ℓ2 +2 +� +n̸=0 +¯αn +n ¯z−n. +(7.50) +The zero-modes are respectively α0 and ¯α0 for ∂X and ¯∂X, and xL and xR for XL and +XR. The meaning of the modes will become clearer in Section 7.1.5 where we study the +commutation relations. +First, we relate the zero-modes α0 and ¯α0 to the conserved charges pL and pR (7.28) of +the U(1) current: +pL = +α0 +√ +2ℓ2 , +pR = +¯α0 +√ +2ℓ2 +(7.51) +such that +X(z) = xL +2 − iℓ2 pL ln z + i +� +ℓ2 +2 +� +n̸=0 +αn +n z−n, +(7.52) +Then, the relations (7.28) and (7.30) allow to rewrite this result in terms of the momentum +p and winding w: +p = +1 +√ +2ℓ2 +� +α0 + ¯α0 +� +, +w = +1 +√ +2ℓ2 +� +α0 − ¯α0 +� +. +(7.53) +These relations can be inverted as +α0 = +� +ℓ2 +2 (p + w), +¯α0 = +� +ℓ2 +2 (p − w). +(7.54) +In the same sense that there are two momenta pL and pR conjugated to xL and xR, +it makes sense to introduce two coordinates x and q conjugated to p and w. From string +theory, the operator x is called the center of mass. The expression (7.54) suggests to write: +xL = x + q, +xR = x − q, +(7.55) +and conversely: +x = 1 +2 (xL + xR), +q = 1 +2 (xL − xR). +(7.56) +In terms of these new variables, the expansion of the full X(z, ¯z) reads: +X(z, ¯z) = x − i ℓ2 +2 +� +p ln |z|2 + w ln z +¯z +� ++ i +� +ℓ2 +2 +� +n̸=0 +1 +n +� +αn z−n + ¯αn ¯z−n� +. +(7.57) +In terms of the coordinates on the cylinder, the part without oscillations becomes: +X(τ, σ) = x − i ℓ2 pτ + ℓ2 wσ + · · · +(7.58) +Note how the presence of ℓ2 gives the correct scale to the second term. The mode q does not +appear at all, and x is the zero-mode of the complete field X(z, ¯z). As it is well-known, the +physical interpretation of x and p is as the position and momentum of the centre-of-mass +of the string.5 If there is a compact dimension, then w counts the number of times the +string winds around it, and q can be understood as the position of the centre-of-mass after +a T-duality.6 +5In worldsheet Lorentzian signature, this becomes X(τ, σ) = x + ℓ2 pt + ℓ2 wσ as expected. +6T-duality and compact bosons fall outside the scope of this book and we refer the reader to [265, +chap. 17, 193, chap. 8] for more details. +114 + +Computation – Equation (7.51) +pL = +1 +2πi +� +dz J = i +ℓ2 +1 +2πi +� +dz ∂X = i +ℓ2 +1 +2πi +� +dz ∂X += +1 +√ +2ℓ2 +1 +2πi +� +dz +� +n +αn z−n−1 = +1 +√ +2ℓ2 α0. +The computation gives pR after replacing α0 by ¯α0. +If the scalar field is non-compact but periodic on the cylinder, the periodicity condition +X(τ, σ + 2π) ∼ X(τ, σ) +(7.59) +translates as +X(e2πiz, e−2πi¯z) ∼ X(z, ¯z). +(7.60) +Evaluating the LHS from (7.50) gives a constraint on the zero-modes: +X(e2πiz, e−2πi¯z) = X(z, ¯z) − i +� +ℓ2 +2 (α0 − ¯α0), +(7.61) +which implies +α0 = ¯α0 +=⇒ +pL = pR = p +2, +w = 0. +(7.62) +The other cases will not be discussed in this book, but we still use the general notation to +make the contact with the literature easier. This also implies that XL and XR cannot be +periodic independently. Hence, the zero-mode couples the holomorphic and anti-holomorphic +sectors together. +The number operators Nn ¯Nn at level n > 0 are defined by: +Nn = ϵ +n α−nαn, +¯Nn = ϵ +n ¯α−n¯αn. +(7.63) +The modes have been normal ordered. They count the number of excitations at the level n: +the factor n−1 is necessary because the modes are not canonically normalized. Then, one +can build the level operators +N = +� +n>0 +n Nn. +(7.64) +They count the number of excitations at level n weighted by the level itself. This corresponds +to the total energy due to the oscillations (the higher the level, the more energy it needs to +be excited). +The Virasoro operators are +Lm = ϵ +2 +� +n +:αnαm−n: +(7.65) +For m ̸= 0, we have +m ̸= 0 : +Lm = ϵ +2 +� +n̸=0,m +:αnαm−n: + ϵ α0αm, +(7.66) +there is no ordering ambiguity and the normal order can be removed. In the case of the +zero-mode, one finds +L0 = ϵ +2 +� +n +:αnα−n: = N + ϵ +2 α2 +0 = N + ϵℓ2 p2 +L, +(7.67) +115 + +using (7.64) and (7.51). It is also useful to define �L0 which corresponds to L0 stripped from +the zero-mode contribution: +�L0 := N. +(7.68) +Similarly, the anti-holomorphic zero-mode is +¯L0 = ¯N + ϵℓ2 p2 +R, +�¯L0 := ¯N, +(7.69) +such that +L+ +0 = N + ¯N + ϵℓ2 (p2 +L + p2 +R) = N + ¯N + ϵℓ2 +2 (p2 + w2), +(7.70a) +L− +0 = N − ¯N + ϵℓ2 (p2 +L − p2 +R) = N − ¯N + ϵℓ2 wp, +(7.70b) +where L± +0 := L0 ± ¯L0 as defined in (6.118). The last equality of each line follows from +(7.31b). The expression of L+ +0 for N = ¯N = 0 matches the weights (7.46) of the vertex +operators for pL = pR = p/2 (no winding), which will be interpreted below. It is a good +place to stress that pL, pR, p and w are operators, while k is a number. +7.1.5 +Commutators +The commutators can be computed from (6.75a) knowing the OPE (7.37). The modes of +∂X and ¯∂X satisfy +[αm, αn] = ϵ m δm+n,0, +[¯αm, ¯αn] = ϵ m δm+n,0, +[αm, ¯αn] = 0 +(7.71) +for all m, n ∈ Z (including the zero-modes). The appearance of the factor m in the RHS +explains the normalization of the number operator (7.63). +From the commutators of the zero-modes, we directly find the ones for the momentum +and winding: +[p, w] = [p, p] = [w, w] = 0, +[p, αn] = [p, ¯αn] = [w, αn] = [w, ¯αn] = 0. +(7.72) +The OPE (7.36) yields +[xL, pL] = iϵ, +[xR, pR] = iϵ, +(7.73) +which can be used to determine the commutators of x and q: +[x, p] = [q, w] = iϵ, +[x, w] = [q, p] = 0. +(7.74) +This shows that (x, p) and (q, w) are pairs of conjugate variables. Interestingly, the winding +number w commutes will all other modes except q, but the latter disappears from the +description. Hence, it can be interpreted as a number which labels different representations: +if no other principle (like periodicity) forbids w ̸= 0, then one can except to have states with +all possible w in the spectrum, each value of w forming a different sector. There are other +interpretations from the point of view of T-duality and double field theory [107, 110, 189, +265]. +The commutator of the modes with the Virasoro operators is +[Lm, αn] = −n αm+n. +(7.75) +as expected from (6.115). For m = 0, this reduces to +[L0, α−n] = n α−n, +(7.76) +which shows that negative modes increase the energy. +The commutator of the creation +modes α−n with the number operators is +[Nm, α−n] = α−mδm,n. +(7.77) +116 + +7.1.6 +Hilbert space +The Hilbert space of the free scalar has the structure of a Fock space. +From (7.76), the momentum p commutes with the Hamiltonian L+ +0 such that it is a +good quantum number to label the states:7 this translates the fact the action (7.1) does not +depend on the conjugate variable x. As a consequence, there exists a family of vacua |k⟩. +The vacua |k⟩ are the states related to the vertex operators (7.33) through the state- +operator correspondence: +|k⟩ := lim +z,¯z→0 Vk(z, ¯z) |0⟩ = eiϵkx |0⟩ , +(7.78) +where |0⟩ is the SL(2, C) vacuum and x is the zero-mode of X(z, ¯z). That this identification +is correct follows by applying the operator p: +p |k⟩ = k |k⟩ . +(7.79) +The notation is consistent with the one of the SL(2, C) vacuum since p |0⟩ = 0. +The vacuum is annihilated by the action of the positive-frequency modes: +∀n > 0 : +αn |k⟩ = 0, +(7.80) +which is equivalent to +Nn |k⟩ = 0. +(7.81) +The different vacua are each ground state of a Fock space (they are all equivalent), but they +are not ground states of the Hamiltonian since they have different energies: +L+ +0 |k⟩ = 2ϵℓ2 k2 |k⟩ , +L− +0 |k⟩ = 0, +(7.82) +using (7.70). The SL(2, C) vacuum is the lowest (highest) energy state if ϵ = 1 (ϵ = −1). +The Fock space F(k) built from the vacuum at momentum k is found by acting repet- +itively with the negative-frequency modes. A convenient basis, the oscillator basis, is given +by the states: +F(k) = Span +� +|k; {Nn}⟩ +� +, +(7.83a) +|k; {Nn}⟩ := +� +n≥1 +(α−n)Nn +� +nNnNn! +|k⟩ , +Nn ∈ N∗ +(7.83b) +(we don’t distinguish the notations between the number operators and their eigenvalues). +The full Hilbert space is given by: +H = +� +R +dk F(k). +(7.84) +Computation – Equation (7.78) +We provide a quick argument to justify the second form of (7.78). Take the limit of +(7.57) with w = 0: +lim +z,¯z→0 eiϵkX(z,¯z) |0⟩ = lim +z,¯z→0 exp iϵk +� +�x − i ℓ2 +2 p ln |z|2 + i +� +ℓ2 +2 +� +n̸=0 +1 +n +� +αn z−n + ¯αn ¯z−n� +� +� |0⟩ += lim +z,¯z→0 exp +� +�iϵkx − ϵk +� +ℓ2 +2 +� +n̸=0 +1 +n +� +αn z−n + ¯αn ¯z−n� +� +� |0⟩ . +7To simplify the discussion, we do not consider winding but only vertex operators of the form (7.33). +117 + +The second term from the first line disappears because p |0⟩ = 0. +For ϵk > 0, as +z, ¯z → 0, the terms with αn and ¯αn for n < 0 disappear since they are accompanied +with a positive power of zn and ¯zn. The modes with n > 0 diverge but the minus sign +makes the exponential to vanish. A more rigorous argument requires to normal order +the exponential and then to use (7.80). +Computation – Equation (7.79) +p |k⟩ = 1 +ℓ2 +1 +2πi +� � +dz i∂X(z) + d¯z i¯∂X(¯z) +� +Vk(0, 0) |0⟩ += 1 +ℓ2 +1 +2πi +� �dz +z +ℓ2k +2 ++ d¯z +¯z +ℓ2k +2 +� +Vk(0, 0) |0⟩ += k Vk(0, 0) |0⟩ +using (7.44). +Remark 7.4 (Fock space and Verma module isomorphism) Note that, in the absence +of the so-called null states, there is a one-to-one map between states in the α−n oscillator +basis and in the L−n Virasoro basis. This translates an isomorphism between the Fock space +and the Verma module of Vk. +One hint for this relation is that applying α−n and L−n +changes the weight (eigenvalue of L0) by the same amount, and there are as many operators +in both basis. +7.1.7 +Euclidean and BPZ conjugates +Since X is a real scalar field, it is self-adjoint (6.95) such that +x† = x +p† = p, +α† +n = α−n. +(7.85) +This implies that the Virasoro operators (7.65) are Hermitian: +L† +n = L−n, +(7.86) +as expected since T(z) is self-adjoint for a free scalar field. +As a consequence of (7.85), the adjoint of the vacuum |k⟩ follows from (7.78): +⟨k| = |k⟩‡ =⟨0| e−iϵkx, +⟨k| p =⟨k| k. +(7.87) +The BPZ conjugate (6.111) of the mode αn is: +αt +n = −(±1)nα−n, +(7.88) +where the sign depends on the choice of I± in (6.111). Using (7.53), this implies that the +momentum operator gets a minus sign:8 +pt = −p, +⟨−k| = |k⟩t . +(7.89) +The inner product between two vacua |k⟩ and |k′⟩ is normalized as: +⟨k|k′⟩ = 2π δ(k − k′) +(7.90) +8Be careful that |k⟩ is not the state associated to the operator p through the state–operator correspond- +ence. Instead, they are associated to Vk, see (7.78). This explains why ⟨k| ̸= (|k⟩)t as in (6.136). +118 + +such that the conjugate state (6.145) of the vacuum reads +⟨kc| = 1 +2π ⟨k| . +(7.91) +The Hermitian and BPZ conjugate states are related as: +|k⟩‡ = − |k⟩t , +(7.92) +which can be interpreted as a reality condition on |k⟩. +7.2 +First-order bc ghost system +First-order systems describe two free fields called ghosts which have a first-order action +and whose conformal weights sum to 1. Commuting (resp. anti-commuting) fields are often +denoted by β and γ (resp. b and c) and correspondingly first-order systems are also called +βγ or bc systems. We will introduce a sign ϵ = ±1 to denote the Grassmann parity of the +fields and always write them as b and c. In string theory, first-order systems describe the +Faddeev–Popov ghosts associated to reparametrizations and supersymmetries (Sections 2.4 +and 17.1). +7.2.1 +Covariant action +A first-order system is defined by two symmetric and traceless fields bµ1···µλ and cµ1···µλ−1 +called ghosts. For fields of integer spins, the dynamics is governed by the first-order action +S = 1 +4π +� +d2x√g gµν bµµ1···µλ−1∇νcµ1···µλ−1 +(7.93) +after taking into account the symmetries of the field indices. Obviously, for λ = 2, one +recovers the reparametrization ghost action (2.145). The action (7.93) is invariant under +Weyl transformations (the fields and covariant derivatives are inert) such that it describes +a CFT on flat space. +When the fields have half-integer spins (and often denoted as β and γ in this case), they +carry a spinor index. In this case, the action contains a Dirac matrix, and the covariant +derivative a spin connection. +The ghost action (7.93) is invariant under a global U(1) symmetry +bµ1···µn −→ e−iθbµ1···µn, +cµ1···µn−1 −→ eiθcµ1···µn−1. +(7.94) +7.2.2 +Action on the complex plane +The simplest description of the system is on the complex plane. Due to the conditions im- +posed on the fields, they have only two independent components for all n, and the equations +of motion imply that one is holomorphic, and the other anti-holomorphic: +b(z) := bz···z(z), +¯b(¯z) := b¯z···¯z(¯z), +c(z) := cz···z(¯z), +¯c(¯z) := c¯z···¯z(z). +(7.95) +In this language, the action becomes +S = 1 +2π +� +d2z +� +b¯∂c + ¯b∂¯c). +(7.96) +This action gives the correct equations of motion +∂¯b = 0, +¯∂b = 0, +∂¯c = 0, +¯∂c = 0. +(7.97) +119 + +Since the fields split into holomorphic and anti-holomorphic sectors, it is convenient to study +only the holomorphic sector as usual. +This system is even simpler than the scalar field +because the zero-modes don’t couple both sectors.9 All formulas for the anti-holomorphic +sector are directly obtained from the holomorphic one by adding bars on quantities, except +for conserved charges which have an index L or R and are both written explicitly. +The action describes a CFT, and the weight of the fields are given by +h(b) = λ, +h(c) = 1 − λ, +h(¯b) = λ, +h(¯c) = 1 − λ, +(7.98) +where λ = n if the fields are in a tensor representation, and λ = n + 1/2 if they are in a +spinor-tensor representation. The holomorphic energy–momentum reads +T = −λ :b∂c: + (1 − λ) :∂b c: +(7.99a) += −λ :∂(bc): + :∂b c: +(7.99b) += (1 − λ) :∂(bc): − :b ∂c:. +(7.99c) +Normal ordering is taken with respect to the SL(2, C) vacuum (6.121). +Finally, both fields can be classically commuting or anticommuting (see below for the +quantum commutators): +b(z)c(w) = −ϵ c(w)b(z), +b(z)b(w) = −ϵ b(w)b(z), +c(z)c(w) = −ϵ c(w)c(z), (7.100) +where ϵ denotes the Grassmann parity +ϵ = +� ++1 +anticommuting, +−1 +commuting. +(7.101) +Sometimes, if ϵ = +1, one denotes b and c respectively by β and γ. If b and c are ghosts +arising from Faddeev–Popov gauge fixing, then ϵ = 1 if λ is integer; and ϵ = −1 if λ is +half-integer (“wrong” spin–statistics assignment). +The U(1) global symmetry (7.94) reads infinitesimally +δb = −ib, +δc = ic, +δ¯b = −i¯b, +δ¯c = i¯c. +(7.102) +It is generated by the conserved ghost current with components: +j(z) = −:b(z)c(z):, +¯ȷ(¯z) = −:¯b(¯z)¯c(¯z): +(7.103) +and the associated charge is called the ghost number +Ngh = Ngh,L + Ngh,R, +Ngh,L = +� +dz +2πi j(z), +Ngh,R = − +� +d¯z +2πi ¯ȷ(¯z). +(7.104) +This charge counts the number of c ghosts minus the number of b ghosts, such that +Ngh(c) = 1, +Ngh(b) = −1, +Ngh(¯c) = 1, +Ngh(¯b) = −1. +(7.105) +The propagator can be derived from the path integral +� +d′b d′c +δ +δb(z) +� +b(w)e−S[b,c]� += 0 +(7.106) +which gives the differential equation +δ(2)(z − w) + 1 +2π ⟨b(w)¯∂c(z)⟩ = 0. +(7.107) +Using (B.2), the solution is easily found to be +⟨c(z)b(w)⟩ = +1 +z − w. +(7.108) +9For the scalar field, the coupling of both sectors happened because of the periodicity condition (7.62). +120 + +Remark 7.5 The propagator is constructed with the path integral. For convenience, the +zero-modes are removed from the measure: reintroducing them, one finds that the propag- +ator is computed not in the conformal vacuum (which has no operator insertion), but in +a state with ghost insertions. This explains why the propagator (7.108) is not of the form +(6.62b). However, this form is sufficient to extract the OPE as changing the vacuum does +not introduce singular terms. +7.2.3 +OPE +The OPEs between the b and c fields are found from the propagator (7.108): +c(z)b(w) ∼ +1 +z − w, +b(z)c(w) ∼ +ϵ +z − w, +(7.109a) +b(z)b(w) ∼ 0, +c(z)c(w) ∼ 0. +(7.109b) +The OPE of each ghost with T confirms the conformal weights in (7.98): +T(z)b(w) ∼ λ +b(w) +(z − w)2 + ∂b(w) +z − w , +(7.110a) +T(z)c(w) ∼ (1 − λ) +c(w) +(z − w)2 + ∂c(w) +z − w . +(7.110b) +The OPE of T with itself is +T(z)T(w) ∼ +cλ/2 +(z − w)4 + +2T(w) +(z − w)2 + ∂T(w) +z − w , +(7.111) +where the central charge is: +cλ = 2ϵ(−1 + 6λ − 6λ2) = −2ϵ +� +1 + 6λ(λ − 1) +� +. +(7.112) +Introducing the ghost charge: +qλ = ϵ(1 − 2λ), +(7.113) +the central charge can also be written as +cλ = ϵ(1 − 3q2 +λ). +(7.114) +This parameter will appear many times in this section and its meaning will become clearer +as we proceed. +The OPE between the ghost current (7.103) and the b and c ghosts read +j(z)b(w) ∼ − b(w) +z − w, +(7.115a) +j(z)c(w) ∼ c(w) +z − w. +(7.115b) +The coefficients of the (z − w)−1 terms correspond to the ghost number of the b and c fields +(7.105). More generally, the ghost number Ngh(O) of any operator O(z) is defined by +j(z)O(w) ∼ Ngh(O) O(w) +z − w. +(7.116) +The OPE for j with itself is +j(z)j(w) ∼ +ϵ +(z − w)2 . +(7.117) +121 + +Finally, the OPE of the current with T reads: +T(z)j(w) ∼ +qλ +(z − w)3 + +j(w) +(z − w)2 + ∂j(w) +z − w . +(7.118) +Due to the presence of the z−3 term, the current j(z) is not a primary field if qλ ̸= 0, that is, +if λ ̸= 1/2. In that case, its transformation under changes of coordinates gets an anomalous +contribution: +j(z) = dw +dz j′(w) + qλ +2 +d +dz ln dw +dz = dw +dz j′(w) + qλ +2 +∂2 +zw +∂zw . +(7.119) +This implies in particular that the currents on the plane and on the cylinder (w = ln z) are +related by: +j(z) = dw +dz +� +jcyl(w) − qλ +2 +� +, +(7.120) +which leads to the following relation between the ghost numbers on the plane and on the +cylinder: +Ngh = N cyl +gh − qλ, +Ngh,L = N cyl +gh,L − qλ +2 , +Ngh,R = N cyl +gh,R − qλ +2 . +(7.121) +For this reason, it is important to make clear the space with respect to which is given the +ghost number: if not explicitly stated, ghost numbers in this book are given on the plane.10 +Due to this anomaly, one finds that the ghost number is not conserved on a curved space: +N c − N b = −ϵ qλ +2 +χg = (1 − 2λ)(g − 1), +(7.122) +where χg is the Euler characteristics (2.4), N b and N c are the numbers of b and c operators. +In string theory, where the only ghost insertions are zero-modes, this translates into a +statement on the number of zero-modes to be inserted. Hence, this can be interpreted as a +generalization of (2.72). For a proof, see for example [24, p. 397]. +Computation – Equation (7.110a) +T(z)b(w) = +� +− λ :b(z)∂c(z): + (1 − λ) :∂b(z) c(z): +� +b(w) +∼ −λ :b(z)∂c(z): b(w) + (1 − λ) :∂b(z) c(z): b(w) +∼ −λ b(z)∂z +1 +z − w + (1 − λ) ∂b(z) +1 +z − w +∼ λ +� +b(w) +(((((( +(z − w)∂b(w) +� +1 +(z − w)2 + (1 − �λ) ∂b(w) +z − w . +10Other references, especially old ones, give it on the cylinder. This can be easily recognized if some ghost +numbers in the holomorphic sector are half-integers: for the reparametrization ghosts, qλ is an integer such +that the shift in (7.121) is a half-integer. +122 + +Computation – Equation (7.110b) +T(z)c(w) = +� +− λ :b(z)∂c(z): + (1 − λ) :∂b(z)c(z): +� +c(w) +∼ ϵλ :∂c(z)b(z): c(w) − ϵ(1 − λ) :c(z)∂b(z): c(w) +∼ λ ∂c(z) +z − w − (1 − λ) c(z) ∂z +1 +z − w +∼ λ ∂c(w) +z − w + (1 − λ) +� +c(w) + (z − w)∂c(w) +� +1 +(z − w)2 +∼ (1 − λ) +c(w) +(z − w)2 + +∂c(w) +(z − w)2 . +Computation – Equation (7.115a) +j(z)b(w) = −:b(z)c(z): b(w) ∼ −:b(z)c(z): b(w) ∼ − b(z) +z − w ∼ − b(w) +z − w. +Computation – Equation (7.115b) +j(z)c(w) = −:b(z)c(z): c(w) ∼ ϵ :c(z)b(z): c(w) ∼ c(z) +z − w ∼ c(w) +z − w. +Computation – Equation (7.117) +j(z)j(w) = :b(z)c(z): :b(w)c(w): +∼ :b(z)c(z): :b(w)c(w): + :b(z)c(z): :b(w)c(w): + :b(z)c(z): :b(w)c(w): +∼ +ϵ +(z − w)2 + ϵ :c(z)b(w): +z − w ++ :b(z)c(w): +z − w +∼ +ϵ +(z − w)2 . +7.2.4 +Mode expansions +The b and c ghosts are expanded as +b(z) = +� +n∈Z+λ+ν +bn +zn+λ , +c(z) = +� +n∈Z+λ+ν +cn +zn+1−λ , +(7.123) +where ν = 0, 1/2 depends on ϵ and on the periodicity of the fields, see (6.102). The modes +are extracted with the contour formulas +bn = +� +dz +2πi zn+λ−1b(z), +cn = +� +dz +2πi zn−λc(z). +(7.124) +Ghosts with λ ∈ Z have integer indices and ν = 0 (we don’t consider ghosts with twisted +boundary conditions). On the other hand, ghosts with λ ∈ Z + 1/2 have integer indices +123 + +and ν = 1/2 in the R sector, and half-integer indices and ν = 0 in the NS sector (see +Section 6.4.4). The choices in the boundary conditions arise from the Z2 symmetry of the +action: +b −→ −b, +c −→ −c. +(7.125) +The number operators N b +n and N c +n are defined to count the numbers of excitations above +the SL(2, C) vacuum of b and c ghosts at level n: +N b +n = :b−ncn:, +N c +n = ϵ :c−nbn:. +(7.126) +The definitions follow from the commutators (7.134). Then, the level operators N b and N c +are obtained by summing over n: +N b = +� +n>0 +n N b +n, +N c = +� +n>0 +n N c +n. +(7.127) +The Virasoro operators are +Lm = +� +n +� +n − (1 − λ)m +� +:bm−ncn: = +� +n +(λm − n) :bncm−n:. +(7.128) +Of particular importance is the zero-mode +L0 = − +� +n +n :bnc−n: = +� +n +n :b−ncn:. +(7.129) +We will give the expression of L0 in terms of the level operators below, see (7.159). To +do this, we will first need to change the normal ordering, which first requires to study the +Hilbert space. +The modes of the ghost current are +jm = − +� +n +:bm−ncn: = − +� +n +:bncm−n:. +(7.130) +Note that the zero-mode of the current also equals the ghost number +Ngh,L = j0 = − +� +n +:b−ncn:. +(7.131) +When both the holomorphic and anti-holomorphic sectors enter, it is convenient to in- +troduce the combinations +b± +n = bn ± ¯bn, +c± +n = 1 +2 (cn ± ¯cn). +(7.132) +The normalization of b± +m is chosen to match the one of L± +m (6.118), and the one of c± +m such +that (7.135) holds. Note the following useful identities: +b− +n b+ +n = 2bn¯bn, +c− +n c+ +n = 1 +2 cn¯cn. +(7.133) +124 + +Computation – Equation (7.128) +T = −λ :b∂c: + (1 − λ) :∂bc: += +� +m,n +� +λ :bmcn: +n + 1 − λ +zm+λzm+2−λ − (1 − λ) :bmcn: +m + λ +zm+λ+1zm+1−λ +� += +� +m,n +� +λ (n + 1 − λ) − (1 − λ)(m + λ) +� :bmcn: +zm−n+2 += +� +m,n +� +λ (n + 1 − λ) − (1 − λ)(m − n + λ) +� :bm−ncn: +zm+2 += +� +m,n +(n − m + λm) :bm−ncn: +zm+2 += +� +m +Lm +zm+2 . +The fourth line follows from shifting m → m−n. The second equality in (7.128) follows +by shifting n → m − n. +Computation – Equation (7.130) +j = −:bc: = +� +m,n +:bmcn: +zm+λzn+1−λ = +� +m,n +:bm−ncn: +zm+1 += +� +m +jm +zm+1 . +7.2.5 +Commutators +The (anti)commutators between the modes bn and cn read: +[bm, cn]ϵ = δm+n,0, +[bm, bn]ϵ = 0, +[cm, cn]ϵ = 0. +(7.134) +Therefore, the modes with n < 0 are creation operators and the modes with n > 0 are +annihilation operators: +• a b ghost excitation at level n > 0 is created by b−n and annihilated by cn; +• a c ghost excitation at level n > 0 is created by c−n and annihilated by bn. +In terms of b± +m and c± +m (7.132), we have: +[b+ +m, c+ +n ]ϵ = δm+n, +[b− +m, c− +n ]ϵ = δm+n. +(7.135) +The commutators of the number operators with the modes are: +[N b +m, b−n] = b−nδm,n, +[N c +m, c−n] = c−nδm,n, +(7.136) +while those between the Ln and the ghost modes are: +[Lm, bn] = +� +m(λ − 1) − n +� +bm+n, +[Lm, cn] = −(mλ + n)cm+n, +(7.137) +in agreement with (6.115). +If n ∈ Z, each ghost field has zero-modes b0 and c0 which +commutes with L0 +[L0, b0] = 0, +[L0, c0] = 0. +(7.138) +125 + +The commutator of the current modes reads +[jm, jn] = m δm+n,0. +(7.139) +Then, the commutator with the Virasoro operators are +[Lm, jn] = −njm+n + qλ +2 m(m + 1)δm+n,0. +(7.140) +Finally, the commutators of the ghost number operator with the ghosts are: +[Ngh, b(w)] = −b(w), +[Ngh, c(w)] = c(w). +(7.141) +Computation – Equation (7.134) +[bm, cn]ϵ = ϵ +� +C0 +dw +2πi w−1 +� +Cw +dz +2πi z−1 wn+λzm−λ+1b(z)c(w) +∼ ϵ +� +C0 +dw +2πi w−1 +� +Cw +dz +2πi z−1 wn+λzm−λ+1 +ϵ +z − w += +� +C0 +dw +2πi wm+n−1 = δm+n,0. +Computation – Equation (7.141) +[Ngh, b(w)] = +� +dz +2πi j(z)b(w) ∼ − +� +dz +2πi +b(w) +z − w = −b(w). +The computation for c is similar. +7.2.6 +Hilbert space +The SL(2, C) vacuum |0⟩ (6.121) is defined by: +∀n > −λ : +bn |0⟩ = 0, +∀n > λ − 1 : +cn |0⟩ = 0. +(7.142) +If λ > 1, there are positive modes which do not annihilate the vacuum. +To simplify the notation, we consider the case λ ∈ Z, the half-integer case following by +shifting the indices by 1/2. Since the modes {c1, . . . , cλ−1} do not annihilate |0⟩, one can +create states +|n1, . . . , nλ−1⟩ = cn1 +1 · · · cnλ−1 +λ−1 |0⟩ +(7.143) +which have negative energies: +L0 |n1, . . . , nλ−1⟩ = − +� +� +λ−1 +� +j=1 +j nj +� +� |n1, . . . , nλ−1⟩ , +(7.144) +where (7.137) has been used. Moreover, this state is degenerate due to the existence of +zero-modes since they commute with the Hamiltonian – see (7.138). As a consequence, it +must be in a representation of the zero-mode algebra. +If the ghosts are commuting (ϵ = −1), then it seems hard to make sense of the theory +since one can find a state of arbitrarily negative energy since ni ∈ N. The zero-modes make +the problem even worse. The appropriate interpretation of these states will be discussed in +the context of the superstring theory for λ = 3/2 (superconformal ghosts). +In the rest of this section, we focus on the Grassmann odd case ϵ = 1. +126 + +Energy vacuum (Grassmann odd) +Since ni = 0 or ni = 1 for anticommuting ghosts (ϵ = 1), there is a state of lowest energy. +This is the energy vacuum (6.129). Since the zero-modes b0 and c0 commute with L0, it is +doubly degenerate. A convenient basis is +� +| ↓⟩ , | ↑⟩ +� +, +(7.145) +where +| ↓⟩ := c1 · · · cλ−1 |0⟩ , +| ↑⟩ := c0c1 · · · cλ−1 |0⟩ . +(7.146) +A general vacuum is a linear combination of the two basis vacua: +|Ω⟩ = ω↓ | ↓⟩ + ω↑ | ↑⟩ , +ω↓, ω↑ ∈ C. +(7.147) +The algebra of these vacua is the one of a two-state system: +b0 | ↑⟩ = | ↓⟩ , +c0 | ↓⟩ = | ↑⟩ , +b0 | ↓⟩ = 0, +c0 | ↑⟩ = 0. +(7.148) +Hence, for the vacuum | ↓⟩ (resp. | ↑⟩), b0 (resp. c0) acts as an annihilation operator, and +conversely c0 (resp. b0) acts as a creation operator. Finally, both states are annihilated by +all positive modes: +∀n > 0 : +bn | ↓⟩ = bn | ↑⟩ = 0, +cn | ↓⟩ = bn | ↓⟩ = 0. +(7.149) +Note that the SL(2, C) vacuum can be recovered by acting with b−n with n < λ: +|0⟩ = b1−λ · · · b−1 | ↓⟩ = b1−λ · · · b−1b0 | ↑⟩ . +(7.150) +The zero-point energy (6.130) of these states is the conformal weight of the vacuum: +L0 | ↓⟩ = aλ | ↓⟩ , +L0 | ↑⟩ = aλ | ↑⟩ , +(7.151) +where aλ can be written in various forms: +aλ = − +λ−1 +� +n=1 +n = −λ(λ − 1) +2 += cλ +24 + 2 +24. +(7.152) +Taking into account the anti-holomorphic sector leads to a four-fold degeneracy. The +basis +� +| ↓↓⟩ , | ↑↓⟩ , | ↓↑⟩ , | ↑↑⟩ +� +, +(7.153) +is built as follows: +| ↓↓⟩ := c1¯c1 · · · cλ−1¯cλ−1 |0⟩ , +| ↑↓⟩ := c0 | ↓↓⟩ , +| ↓↑⟩ := ¯c0 | ↓↓⟩ , +| ↑↑⟩ := c0¯c0 | ↓↓⟩ . +(7.154) +The modes b0 and ¯b0 can be used to flip the arrows downward, leading to the following +algebra: +c0 | ↓↓⟩ = | ↑↓⟩ , +¯c0 | ↓↓⟩ = | ↓↑⟩ , +c0 | ↓↑⟩ = −¯c0 | ↑↓⟩ = | ↑↑⟩ , +b0 | ↑↑⟩ = | ↓↑⟩ , +¯b0 | ↑↑⟩ = − | ↑↓⟩ , +b0 | ↑↓⟩ = ¯b0 | ↓↑⟩ = | ↓↓⟩ , +(7.155a) +The vacua are annihilated by different combinations of the zero-modes: +b0 | ↓↓⟩ = ¯b0 | ↓↓⟩ = 0, +c0 | ↑↓⟩ = ¯b0 | ↑↓⟩ = 0, +b0 | ↓↑⟩ = ¯c0 | ↓↑⟩ = 0, +c0 | ↑↑⟩ = ¯c0 | ↑↑⟩ = 0. +(7.155b) +127 + +In these manipulations, one has to be careful to correctly anti-commute the modes with the +ones hidden in the definitions of the vacua. +There is a second basis which is more natural when using the zero-modes c± +0 and b± +0 +(7.132): +� +| ↓↓⟩ , |+⟩ , |−⟩ , | ↑↑⟩ +� +, +(7.156) +where the two vacua |±⟩ are combinations of the | ↓↑⟩ and | ↑↓⟩ vacua: +|±⟩ = | ↑↓⟩ ± | ↓↑⟩ . +(7.157) +The different vacua are naturally related by acting with c± +0 and b± +0 which act as raising and +lowering operators: +c± +0 | ↓↓⟩ = 1 +2 |±⟩ , +c∓ +0 |±⟩ = ± | ↑↑⟩ , +b± +0 |±⟩ = ±2 | ↓↓⟩ , +b∓ +0 | ↑↑⟩ = ± |±⟩ . +(7.158a) +From the previous relations, it follows that the different vacua are annihilated by the zero- +modes as follow: +b+ +0 | ↓↓⟩ = b− +0 | ↓↓⟩ = 0, +c− +0 |−⟩ = b+ +0 |−⟩ = 0, +c+ +0 |+⟩ = b− +0 |+⟩ = 0 +c+ +0 | ↑↑⟩ = c− +0 | ↑↑⟩ = 0, +(7.158b) +This also means that we have +c− +0 c+ +0 | ↓↓⟩ = 1 +2 | ↑↑⟩ , +b+ +0 b− +0 | ↑↑⟩ = 2 | ↓↓⟩ . +(7.158c) +Computation – Equation (7.158) +2 c+ +0 |±⟩ = (c0 + ¯c0) | ↑↓⟩ ± (c0 + ¯c0) | ↓↑⟩ = ¯c0 | ↑↓⟩ ± c0 | ↓↑⟩ = (−1 ± 1) | ↑↑⟩ +b+ +0 |±⟩ = (b0 + ¯b0) | ↑↓⟩ ± (b0 + ¯b0) | ↓↑⟩ = b0 | ↑↓⟩ ± ¯b0 | ↓↑⟩ = (1 ± 1) | ↓↓⟩ +2 c± +0 | ↓↓⟩ = (c0 ± ¯c0) | ↓↓⟩ = c0 | ↓↓⟩ ± ¯c0 | ↓↓⟩ = | ↑↓⟩ ± | ↓↑⟩ = |±⟩ +b± +0 | ↑↑⟩ = (b0 ± ¯b0) | ↑↑⟩ = b0 | ↑↑⟩ ± ¯b0 | ↑↑⟩ = | ↓↑⟩ ∓ | ↑↓⟩ = ∓ |∓⟩ +Energy normal ordering (Grassmann odd) +We now turn towards the definition of the energy normal ordering (6.150). Ultimately, it +will be found that | ↓⟩ is the physical vacuum in string theory. For this reason, the energy +normal ordering +⋆ +⋆ · · · +⋆ +⋆ is associated to the vacuum | ↓⟩ in order to resolve the ambiguity +of the zero-modes. In particular, b0 is an annihilation operator in this case, while c0 is a +creation operator. In the rest of this section, we translate the normal ordering of expressions +from the conformal vacuum to the energy vacuum. +The Virasoro operators Ln for n ̸= 0 have no ordering problems since the modes which +compose them commute. The expression of L0 (7.129) in the energy ordering becomes +L0 = +� +n +n +⋆ +⋆b−ncn +⋆ +⋆ + aλ = N b + N c + aλ +(7.159) +where aλ is the zero-point energy (7.152) and N b and N c are the ghost mode numbers +(7.127). The contribution of the non-zero modes is denoted by: +�L0 = N b + N c. +(7.160) +128 + +The expression can be rewritten to encompass all modes: +Lm = +� +n +� +n − (1 − λ)m +� ⋆ +⋆bm−ncn +⋆ +⋆ + aλ δm,0 +(7.161) +Similarly, the expression of the ghost number is +Ngh,L = j0 = +� +n +⋆ +⋆b−ncn +⋆ +⋆ − +�qλ +2 + 1 +2 +� +(7.162a) += +� +n>0 +� +N c +n − N b +n +� ++ 1 +2 +� +N c +0 − N b +0 +� +− qλ +2 , +(7.162b) +and thus: +jm = +� +n +⋆ +⋆bm−ncn +⋆ +⋆ − +�qλ +2 + 1 +2 +� +δm,0. +(7.163) +It is useful to define the ghost number without ghost zero-modes: +� +Ngh,L := +� +n>0 +� +N c +n − N b +n +� +. +(7.164) +One can straightforwardly compute the ghost number of the vacua: +j0 | ↓⟩ = (λ − 1) | ↓⟩ = +� +−qλ +2 − 1 +2 +� +| ↓⟩ , +(7.165a) +j0 | ↑⟩ = λ | ↑⟩ = +� +−qλ +2 + 1 +2 +� +| ↑⟩ . +(7.165b) +This confirms that the SL(2, C) vacuum has vanishing ghost number since | ↓⟩ contains +exactly λ − 1 ghosts: +j0 |0⟩ = 0. +(7.166) +Using (7.121) allows to write the ghost numbers on the cylinder: +jcyl +0 +| ↓⟩ = −1 +2 | ↓⟩ , +jcyl +0 +| ↑⟩ = 1 +2 | ↑⟩ . +(7.167) +That both ghost numbers have same magnitude but opposite signs could be expected: since +the ghost number changes as Ngh → −Ngh when b ↔ c, the mean value of the ghost number +should be zero. +Remark 7.6 (Ghost number conventions) Since the ghost number is an additive quan- +tum number, it is always possible to shift its definition by a constant. This can be used to +set the ghost numbers of the vacua to some other values. For example, [24, p. 116] adds +qλ/2 to the ghost number in order to get Ngh = ±1/2 on the plane (instead of the cylinder). +We do not follow this convention in order to keep the symmetry between the vacuum ghost +numbers on the cylinder. +129 + +Computation – Equation (7.159) +Start with (7.129) and use (6.157): +L0 = − +� +n +n :bnc−n: = − +� +n≤−λ +n bnc−n + ϵ +� +n>−λ +n c−nbn += +� +n≥λ +n b−ncn + ϵ +� +n>−λ +n c−nbn += +� +n≥λ +n b−ncn + ϵ +� +n>0 +n c−nbn + ϵ +0 +� +n=−λ+1 +n c−nbn += +� +n≥λ +n b−ncn + ϵ +� +n>0 +n c−nbn + ϵ +λ−1 +� +n=0 +n b−ncn + aλ += +� +n>0 +n b−ncn + ϵ +� +n>0 +n c−nbn + aλ, += +� +n +n +⋆ +⋆b−ncn +⋆ +⋆ + aλ, +using that +0 +� +n=−λ+1 +c−nbn = − +λ−1 +� +n=0 +n cnb−n = − +λ−1 +� +n=0 +n (−ϵ b−ncn + 1) = ϵ +λ−1 +� +n=0 +n b−ncn + aλ. +The result also follows from (6.163). +Computation – Equation (7.162) +j0 = − +� +n +:b−ncn: = − +� +n≥λ +b−ncn + ϵ +� +n>−λ +c−nbn += − +� +n≥λ +b−ncn + ϵ +� +n>0 +c−nbn + ϵ +λ−1 +� +n=1 +cnb−n + ϵ c0b0 += − +� +n≥λ +b−ncn + ϵ +� +n>0 +c−nbn − +λ−1 +� +n=1 +b−ncn + ϵ(λ − 1) + ϵ c0b0 += − +� +n>0 +b−ncn + ϵ +� +n>0 +c−nbn + ϵ(λ − 1) + ϵ c0b0. +Finally, one can write +ϵ(λ − 1) = −qλ +2 − ϵ +2. +(7.168) +The result also follows from (6.163). The second expression is obtained by symmetrizing +the last term such that +ϵ c0b0 + ϵ(λ − 1) = ϵ +2 c0b0 + 1 +2(−b0c0 + ϵ) + ϵ(λ − 1) += 1 +2 (ϵ c0b0 − b0c0) + ϵ +� +λ − 1 +2 +� +. +130 + +Structure of the Hilbert space (Grassmann odd) +Since the zero-modes commute with the Hamiltonian and with all other negative- and +positive-frequency modes, the Hilbert space is decomposed in several subspaces, each as- +sociated to a zero-mode.11 +Starting with the holomorphic sector only, the Hilbert space Hgh is: +Hgh = Hgh,0 ⊕ c0Hgh,0, +Hgh,0 := Hgh ∩ ker b0, +(7.169) +which follows from the 2-state algebra (7.148). Obviously, one has c0Hgh,0 = Hgh ∩ ker c0. +The oscillator basis of the Hilbert space Hgh,0 is generated by applying the negative- +frequency modes and has the structure of a fermionic Fock space without zero-modes: +Hgh,0 = Span +� ��↓; {N b +n}; {N c +n} +� � +, +(7.170a) +��↓; {N b +n}; {N c +n} +� += +� +n≥1 +(b−n)Nb +n(c−n)Nc +n | ↓⟩ , +N b +n, N c +n ∈ N∗ +(7.170b) +(again, number operators and their eigenvalues are not distinguished). This means that +Hgh,0 can also be regarded as a Fock space built on the vacuum | ↓⟩, for which c0 and b0 +are respectively creation and annihilation operators. Conversely, c0 and b0 are respectively +annihilation and creation operators for c0Hgh,0. +In particular, this means that any state can be written as the sum of two states +ψ = ψ↓ + ψ↑, +ψ↓ ∈ Hgh,0, +ψ↑ ∈ c0Hgh,0, +(7.171) +with ψ↓ and ψ↑ built respectively on top of the | ↓⟩ and | ↑⟩ vacua. +This pattern generalizes when considering both the holomorphic and anti-holomorphic +sectors. In that case, the Hilbert space is decomposed in four subspaces:12 +Hgh = Hgh,0 ⊕ c0Hgh,0 ⊕ ¯c0Hgh,0 ⊕ c0¯c0Hgh,0, +Hgh,0 := Hgh ∩ ker b0 ∩ ker¯b0. +(7.172) +Basis states of the Hilbert space Hgh,0 are: +��↓↓; {N b +n}; {N c +n}; { ¯N b +n}; { ¯N c +n} +� += +� +n≥1 +(b−n)N b +n(¯b−n) +¯ +Nb +n(c−n)Nc +n(¯c−n) +¯ +Nc +n | ↓↓⟩ , +N b +n, ¯N b +n, N c +n, ¯N c +n ∈ N∗. +(7.173) +A general state of Hgh can be decomposed as +ψ = ψ↓↓ + ψ↑↓ + ψ↓↑ + ψ↑↑, +(7.174) +where each state is built by acting with negative-frequency modes on the corresponding +vacuum. +In terms of the second basis (7.156), the Hilbert space admits a second decomposition: +Hgh = Hgh,0 ⊕ c+ +0 Hgh,0 ⊕ c− +0 Hgh,0 ⊕ c− +0 c+ +0 Hgh,0, +Hgh,0 := Hgh ∩ ker b− +0 ∩ ker b+ +0 . +(7.175) +11Due to the specific structure of the inner product defined below, these subspaces are not orthonormal +to each other. +12The reader should not get confused by the same symbol Hgh,0 as in the case of the holomorphic sector. +131 + +In view of applications to string theory, it is useful to introduce two more subspaces: +Hgh,± := Hgh ∩ ker b± +0 = Hgh,0 ⊕ c∓ +0 Hgh,0, +(7.176) +and the associated decomposition +Hgh = Hgh,± ⊕ c± +0 Hgh,±. +(7.177) +In off-shell closed string theory, the principal Hilbert space will be H− +gh due to the level- +matching condition. In this case, H− +gh has the same structure as Hgh in the pure holomorphic +sector, and c+ +0 plays the same role as c0. A state in H− +gh is built on top of the vacua | ↓↓⟩ +and |+⟩. +7.2.7 +Euclidean and BPZ conjugates +In order for the Virasoro operators to be Hermitian, the bn and cn must satisfy the following +conditions: +b† +n = ϵb−n, +c† +n = c−n. +(7.178) +Hence, bn is anti-Hermitian if ϵ = −1. The BPZ conjugates of the modes are: +bt +n = (−1)λ b−n, +ct +n = (−1)1−λ c−n, +(7.179) +using I+(z) with (6.111). +In the rest of this section, we consider only the case ϵ = 1 and λ ∈ N. The adjoints of +the vacuum read: +| ↓⟩‡ =⟨0| c1−λ · · · c−1, +| ↑⟩‡ =⟨0| c1−λ · · · c−1c0. +(7.180) +The BPZ conjugates of the vacua are: +⟨↓ | := | ↓⟩t = (−1)(1−λ)2⟨0| c−1 · · · c1−λ, +⟨↑ | := | ↑⟩t = (−1)λ(1−λ)⟨0| c0c−1 · · · c1−λ. +(7.181) +The signs are inconvenient but will disappear when considering both the left and right vacua +together as in (7.154). We have the following relations: +⟨↓ | = (−1)aλ+(1−λ)(2−λ) | ↓⟩‡ , +⟨↑ | = (−1)aλ | ↑⟩‡ , +(7.182) +where aλ is the zero-point energy (7.152). +Computation – Equation (7.182) +To prove the relation, we can start from the BPZ conjugate ⟨↓ | and reorder the modes +to bring them in the same order as the adjoint: +⟨↓ | = (−1)(1−λ)2+ 1 +2 (2−λ)(1−λ) | ↓⟩‡ = (−1)−aλ+(1−λ)(2−λ) | ↓⟩‡ +The reordering gives a factor (−1) to the power: +λ−2 +� +i=1 +i = 1 +2(2 − λ)(1 − λ) = −aλ + 1 − λ. +Similarly, for the second vacuum: +⟨↑ | = (−1)λ(1−λ)− 1 +2 λ(1−λ) | ↑⟩‡ = (−1) +1 +2 λ(1−λ) | ↑⟩‡ . +132 + +We can identify the power with (7.152). +Then, we have the following relations: +⟨↑ | b0 =⟨↓ | , +⟨↓ | c0 =⟨↑ | , +⟨↓ | b0 = 0, +⟨↑ | c0 = 0. +(7.183) +There is a subtlety in defining the inner product because the vacuum is degenerate. If +we write the two vacua as vectors +| ↓⟩ = +� +0 +1 +� +, +| ↑⟩ = +� +1 +0 +� +, +(7.184) +then the zero-modes have the following matrix representation: +b0 = +�0 +0 +1 +0 +� +, +c0 = +�0 +1 +0 +0 +� +. +(7.185) +These matrices are not Hermitian as required by (7.178): since Hermiticity follows from the +choice of an inner product, it means that the vacua cannot form an orthonormal basis. An +appropriate choice for the inner products is:13 +⟨↓ | ↓⟩ = ⟨↑ | ↑⟩ = 0, +⟨↑ | ↓⟩ =⟨↓ | c0 | ↓⟩ =⟨0| c1−λ · · · c−1c0c1 · · · cλ−1 |0⟩ = 1. +(7.186) +The effect of changing the definition of the inner product or to consider a non-orthonormal +basis is represented by the insertion of c0. The last condition implies that the conjugate +state (6.145) to the SL(2, C) vacuum is: +⟨0c| =⟨0| c1−λ · · · c−1c0c1 · · · cλ−1, +⟨↓c | =⟨↑ | . +(7.187) +7.2.8 +Summary +In this section we summarize the values of the parameters for different theories of interest +(Table 7.1). The (η, ξ) system will be introduced in Chapter 17 in the bosonization of the +super-reparametrization (β, γ) ghosts. +The ψ± system can be used to describe spin-1/2 +fermions. +ϵ +λ +qλ +cλ +aλ +b, c (diff.) +1 +2 +−3 +−26 +−1 +β, γ (susy.) +−1 +3/2 +2 +11 +3/8 +ψ± +1 +1/2 +0 +1 +0 +η, ξ +1 +1 +−1 +−2 +0 +Table 7.1: Summary of the first-order systems. Remember that h(b) = λ and h(c) = 1 − λ. +7.3 +Suggested readings +• Free scalar: general references [246, sec. 4.1.3, 4.3, 4.6.2, 54, sec. 5.3.1, 6.3, 24, sec. 4.2, +193, 128], topological current and winding [109, 265, sec. 17.2–3]. +• First-order system: general references [24, chap. 5, sec. 13.1, 128, sec. 4.15, 193, +sec. 2.5], ghost vacua [151, sec. 15.3]. +13To avoid confusions, let us note that the adjoint in (7.182) are defined only through the adjoint of +the modes (6.110) but not with respect to the inner product given here, which would lead to exchanging +| ↓⟩‡ ∼⟨↑ | and | ↑⟩‡ ∼⟨↓ |. +133 + +Chapter 8 +BRST quantization +Abstract +The BRST quantization can be introduced either by following the standard +QFT treatment (outlined in Section 3.2), or by translating it in the CFT language. One +can then use all the CFT techniques to extract information on the spectrum, which makes +this approach more powerful. Moreover, this also provides an elegant description of states +and string fields. In this chapter, we set the stage of the BRST quantization using the CFT +language and we apply it to string theory. The main results of this chapter are a proof of +the no-ghost theorem and a characterization of the BRST cohomology (physical states). +8.1 +BRST for reparametrization invariance +The BRST symmetry we are interested in results from gauge fixing the reparametrization +invariance. +In this chapter, we focus on the holomorphic sector: since both sectors are +independent, most results follow directly, except those concerning the zero-modes. +We +consider a generic matter CFT coupled to reparametrization ghosts: +1. matter: central charge cm, energy–momentum tensor Tm and Hilbert space Hm; +2. reparametrization ghosts: bc ghost system (Sections 2.3 and 7.2) with ϵ = +1 and +λ = 2, cgh = −26, energy–momentum tensor T gh and Hilbert space Hgh. +The formulas for the reparametrization ghosts are summarized in Appendix B.3.5. +For +modes, the system (m, gh, b or c) is indicated as a superscript to not confuse it with the +mode index. +The total central charge, energy–momentum tensor and Hilbert space are +denoted by: +c = cm + cgh = cm − 26, +T(z) = T m(z) + T gh(z), +H = Hm ⊗ Hgh. +(8.1) +The goal is to find the physical states in the cohomology, that is, which are BRST closed +QB |ψ⟩ = 0 +(8.2) +but non exact (Section 3.2): the latter statement can be understood as an equivalence +between closed states under shift by exact states: +|ψ⟩ ∼ |ψ⟩ + QB |Λ⟩ . +(8.3) +We introduce the BRST current and study its CFT properties. Then, we give a compu- +tation of the BRST cohomology when the matter CFT contains at least two scalar fields. +134 + +8.2 +BRST in the CFT formalism +The BRST current can be found from (3.50) to be [193]: +jB(z) = :c(z) +� +T m(z) + 1 +2 T gh(z) +� +: + κ ∂2c(z) +(8.4a) += c(z)T m(z) + :b(z)c(z)∂c(z): + κ ∂2c(z), +(8.4b) +and similarly for the anti-holomorphic sector. This can be derived from (3.53): the generator +of infinitesimal changes of coordinates (given by the Lie derivative) is the energy–momentum +tensor. The factor of 1/2 comes from the expression (7.99) of the ghost energy–momentum +tensor: the second term does not contribute while the first has a factor of 2. Since the +transformation of c in (3.53) has no factor, the 1/2 is necessary to recover the correct +normalization. Finally, one finds that the transformation of b is reproduced. The different +computations can be checked using the OPEs given below. The last piece is a total derivative +and does not contribute to the charge: for this reason, it cannot be derived from (3.53), its +coefficient will be determined below. Note that it is the only total derivative of dimension +1 and of ghost number 1. +The BRST charge is then obtained by the contour integral: +QB = QB,L + QB,R, +QB,L = +� +dz +2πi jB(z), +QB,R = +� +d¯z +2πi ¯ȷB(¯z). +(8.5) +As usual, QB ∼ QB,L when considering only the holomorphic sectors such that we generally +omit the index. +8.2.1 +OPE +The OPE of the BRST current with T is +T(z)jB(w) ∼ +�cm +2 − 4 − 6κ +� +c(w) +(z − w)4 + (3 − 2κ) ∂c(w) +(z − w)3 + +jB(w) +(z − w)2 + ∂jB(w) +z − w . +(8.6) +Hence, the BRST current is a primary operator only if +cm = 26, +κ = 3 +2. +(8.7) +The BRST current must be primary, otherwise, the BRST symmetry is anomalous, which +means that the theory is not consistent. This provides another derivation of the critical +dimension. In this case, the OPE becomes +T(z)jB(w) ∼ +jB(w) +(z − w)2 + ∂jB(w) +z − w . +(8.8) +Remark 8.1 (Critical dimension in 2d gravity) The value cm = 26 (critical dimen- +sion) was obtained in Section 2.3 by requiring that the Liouville field decouples from the +path integral. In 2d gravity, where this condition is not necessary, (nor even desirable) the +Liouville field is effectively part of the matter, such that cL + cm = 26. One can also study +the BRST cohomology in this case. +The OPE of jB(z) with the ghosts are +jB(z)b(w) ∼ +2κ +(z − w)3 + +j(w) +(z − w)2 + T(w) +z − w, +(8.9a) +jB(z)c(w) ∼ :c(w)∂c(w): +z − w +. +(8.9b) +135 + +Similarly, the OPE with any matter weight h primary field φ is +jB(z)φ(w) ∼ h c(w)φ(w) +(z − w)2 + :h ∂c(w)φ(w) + c(w)∂φ(w): +z − w +, +(8.9c) +using that c(w)2 = 0 to cancel one term. +The OPE with the ghost current is +jB(z)j(w) ∼ +2κ + 1 +(z − w)3 − 2∂c(w) +(z − w)2 − jB(w) +z − w , +(8.10) +while the OPE with itself is (for κ = 3/2) +jB(z)jB(w) ∼ −cm − 18 +2 +:c(w)∂c(w): +(z − w)3 +− cm − 18 +4 +:c(w)∂2c(w): +(z − w)2 +− cm − 26 +12 +:c(w)∂3c(w): +z − w +. +(8.11) +There is no first order pole if cm = 26: as we will see shortly, this implies that the BRST +charge is nilpotent. +8.2.2 +Mode expansions +The mode expansion of the BRST charge can be written equivalently +QB = +� +m +:cm +� +Lm +−m + 1 +2 Lgh +−m +� +: +(8.12a) += +� +m +c−mLm +m + 1 +2 +� +m,n +(n − m) :c−mc−nbm+n: +(8.12b) +In the energy ordering, this expression becomes +QB = +� +m +⋆ +⋆cm +� +Lm +−m + 1 +2 Lgh +−m +� +⋆ +⋆ − c0 +2 +(8.13a) += +� +n +cmLm +−m + 1 +2 +� +m,n +(n − m) +⋆ +⋆c−mc−nbm+n +⋆ +⋆ − c0, +(8.13b) +where the ordering constant is the same as in Lgh +0 +(as can be checked by comparing both +sides of the anticommutator). The simplest derivation of this term is to use the algebra +and to ensure that it is consistent. The only ambiguity is in the second term, when one c +does not commute with the b: this happens for −n + (m + n) = 0, such that the ordering +ambiguity is proportional to c0. Then, one finds that it is equal to agh = −1. +The BRST operator can be decomposed on the ghost zero-modes as +QB = c0L0 − b0M + �QB +(8.14a) +where +�QB = +� +m̸=0 +c−mLm +m − 1 +2 +� +m,n̸=0 +m+n̸=0 +(m − n) +⋆ +⋆c−mc−nbm+n +⋆ +⋆ , +(8.14b) +M = +� +m̸=0 +m c−mcm +(8.14c) +136 + +The interest of this decomposition is that L0, M and �Q do not contain b0 or c0, which +make it very useful to act on states decomposed according to the zero-modes (7.169). The +nilpotency of the BRST operator implies the relations +[L0, M] = [ �QB, M] = [ �QB, L0] = 0, +�Q2 +B = L0M. +(8.15) +Moreover, one has Ngh( �QB) = 1 and Ngh(M) = 2. +8.2.3 +Commutators +From the various OPEs, one can compute the (anti-)commutators of the BRST charge with +the other operators. For the ghosts and a weight h primary field φ, one finds +{QB, b(z)} = T(z), +(8.16a) +{QB, c(z)} = c(z)∂c(z), +(8.16b) +[QB, φ(z)] = h ∂c(z)φ(z) + c(z)∂φ(z). +(8.16c) +This reproduces correctly (3.53). +Two facts will be useful in string theory. First, (8.16c) is a total derivative for h = 1: +[QB, φ(z)] = ∂ +� +c(z)φ(z) +� +. +(8.17) +Second, c(z)φ(z) is closed if h = 1 +{QB, c(z)φ(z)} = (1 − h)c(z)∂c(z)φ(z). +(8.18) +The commutator with the ghost current is +[QB, j(z)] = −jB(z), +(8.19) +which confirms that the BRST charge increases the ghost number by 1 +[Ngh, QB] = QB. +(8.20) +One finds that the BRST charge is nilpotent +{QB, QB} = 0 +(8.21) +and commutes with the energy–momentum tensor +[QB, T(z)] = 0 +(8.22) +only if the matter central charge corresponds to the critical dimension: +cm = 26. +(8.23) +The most important commutator for the modes is +Ln = {QB, bn}. +(8.24) +Nilpotency of QB then implies that QB commutes with Ln: +[QB, Ln] = 0. +(8.25) +137 + +8.3 +BRST cohomology: two flat directions +The simplest case for studying the BRST cohomology is when the target spacetime has at +least two non-compact flat directions represented by two free scalar fields (X0, X1) (Sec- +tion 7.1). The remaining matter fields are arbitrary as long as the critical dimension cm = 26 +is reached. The reason for introducing two flat directions is that the cohomology is easily +worked out by introducing light-cone (or complex) coordinates in target spacetime. +The field X0 can be spacelike or timelike ϵ0 = ±1, while we consider X1 to be always +spacelike, ϵ1 = 1. The oscillators are denoted by α0 +m and α1 +m, and the momenta of the Fock +vacua by k∥ = (k0, k1) such that +k2 +∥ = ϵ0(k0)2 + (k1)2. +(8.26) +The rest of the matter sector, called the transverse sector ⊥, is an arbitrary CFT with +energy–momentum tensor T ⊥, central charge c⊥ = 24 and Hilbert space H⊥. The ghost +together with the two scalar fields form the longitudinal sector ∥. The motivation for the +names longitudinal and transverse will become clear later: they will be identified with the +light-cone and perpendicular directions in the target spacetime (and, correspondingly, with +unphysical and physical states). +The Hilbert space of the theory is decomposed as +H := H∥ ⊗ H⊥, +H∥ := +� +dk0 F0(k0) ⊗ +� +dk1 F1(k1) ⊗ Hgh, +(8.27) +where F0(k0) and F1(k1) are the Fock spaces (7.83a) of the scalar fields X0 and X1, and +Hgh is the ghost Hilbert space (7.169). As a consequence, a generic state of H reads +|ψ⟩ = |ψ∥⟩ ⊗ |ψ⊥⟩ , +(8.28) +where ψ⊥ is a generic state of the transverse matter CFT H⊥ and ψ∥ is built by acting with +oscillators on the Fock vacuum of H∥: +|ψ∥⟩ = cNc +0 +0 +� +m>0 +(α0 +−m)N0 +m(α1 +−m)N1 +m (b−m)Nb +m(c−m)Nc +m |k0, k1, ↓⟩ +|k0, k1, ↓⟩ := |k0⟩ ⊗ |k1⟩ ⊗ | ↓⟩ , +N 0 +m, N 1 +m ∈ N, +N b +m, N c +m = 0, 1. +(8.29) +Since the Virasoro modes commute with the ghost number, eigenstates of the Virasoro +operators without zero-modes �L0, given by the sum of (7.68) and (7.160), can also be taken +to be eigenstates of Ngh. It is also useful to define the Hilbert space of states lying in the +kernel of b0: +H0 = H ∩ ker b0 +(8.30) +such that +H = H0 ⊕ c0H0. +(8.31) +The full L0 operator reads +L0 = Lm +0 + Lgh +0 = (Lm +0 − 1) + N b + N c, +(8.32) +using (7.129) for Lgh +0 . A more useful expression is obtained by separating the two sectors +and by extracting the zero-modes using (7.67): +L0 = +� +L⊥ +0 − m2 +∥,Lℓ2 − 1 +� ++ �L∥ +0, +(8.33) +where the longitudinal mass and total level operator are: +m2 +∥,L = −p2 +∥,L, +�L∥ +0 = N 0 + N 1 + N b + N c ∈ N. +(8.34) +138 + +A state |ψ⟩ is said to be on-shell if it is annihilated by L0: +on-shell: +L0 |ψ⟩ = 0. +(8.35) +The absolute BRST cohomology Habs(QB) defines the physical states (Section 3.2) and +is given by the states ψ ∈ H that are QB-closed but not exact: +Habs(QB) := +� +|ψ⟩ ∈ H +�� QB |ψ⟩ = 0, ∄ |χ⟩ ∈ H +�� |ψ⟩ = QB |χ⟩ +� +. +(8.36) +Since QB commutes with L0, (8.25), the cohomology subspace is preserved under time +evolution. +Before continuing, it is useful to outline the general strategy for studying the cohomology +of a BRST operator Q in the CFT language. The idea is to find an operator ∆ – called +contracting homotopy operator – which, if it exists, trivializes the cohomology. Conversely, +this implies that the cohomology is to be found within states which are annihilated by ∆ +or for which ∆ is not defined. Then, it is possible to restrict Q on these subspaces: this is +advantageous when the restriction of the BRST charge on these subspaces is a simpler. In +fact, we will find that the reduced operator is itself a BRST operator, for which one can +search for another contracting homotopy operator.1 +Given a BRST operator Q, a contracting homotopy operator ∆ for Q is an operator such +that +{Q, ∆} = 1. +(8.37) +Interpreting Q as a derivative operator, ∆ corresponds to the Green function or propagator. +The existence of a well-defined ∆ with empty kernel implies that the cohomology is empty +because all closed states are exact. Indeed, consider a state |ψ⟩ ∈ H which is an eigenstate +of ∆ and closed QB |ψ⟩ = 0. Inserting (8.37) in front of the state gives: +|ψ⟩ = {QB, ∆} |ψ⟩ = QB +� +∆ |ψ⟩ +� +. +(8.38) +If ∆ is well-defined on |ψ⟩ and |ψ⟩ /∈ ker ∆, then ∆ |ψ⟩ is another state in H, which implies +that |ψ⟩ is exact. Hence, the BRST cohomology has to be found inside the subspaces ker ∆ +or on which ∆ is not defined. +8.3.1 +Conditions on the states +In this subsection, we apply explicitly the strategy just discussed to get conditions on the +states. A candidate contracting homotopy operator for QB is +∆ := b0 +L0 +(8.39) +thanks to (8.24): +L0 = {QB, b0}. +(8.40) +Indeed, suppose that |ψ⟩ is an eigenstate of L0, and that it is closed but not on-shell: +QB |ψ⟩ = 0, +L0 |ψ⟩ ̸= 0. +(8.41) +One can use (8.40) in order to write: +|ψ⟩ = QB +� b0 +L0 +|ψ⟩ +� +. +(8.42) +1A similar strategy shows that there is no open string excitation for the open SFT in the tachyon vacuum. +139 + +The operator inside the parenthesis is ∆ defined above in (8.39). The formula (8.42) breaks +down if ψ is in the kernel of L0 since the inverse is not defined. This implies that a necessary +condition for a L0-eigenstate |ψ⟩ to be in the BRST cohomology is to be on-shell (8.35). +Considering explicitly the subset of states annihilated by b0 is not needed at this stage since +ker b0 ⊂ ker L0 for QB-closed states, according to (8.24). Hence, we conclude: +Habs(QB) ⊂ ker L0. +(8.43) +Note that this statement holds only at the level of vector spaces, i.e. when considering +equivalence classes of states |ψ⟩ ∼ |ψ⟩+Q |Λ⟩. This means that there exists a representative +state of each equivalence class inside ker L0, but a generic state is not necessarily in ker L0. +For example, consider a state |ψ⟩ ∈ ker L0 and closed. Then, |ψ′⟩ = |ψ⟩ + QB |Λ⟩ with +|Λ⟩ /∈ ker L0 is still in Habs(QB) but |ψ′⟩ /∈ ker L0 since [L0, QB] = 0. +Computation – Equation (8.42) +For L0 |ψ⟩ ̸= 0, one has: +|ψ⟩ = L0 +L0 +|ψ⟩ = 1 +L0 +{QB, b0} |ψ⟩ = 1 +L0 +QB +� +b0 |ψ⟩ +� +where the fact that |ψ⟩ is closed has been used to cancel the second term of the anti- +commutator. Note that L0 commutes with both QB and b0 such that it can be moved +freely. +This shows that ∆ = b0/L0 given by (8.39) is not a contracting homotopy operator. A +proper definition involves the projector P0 on the kernel of L0: +|ψ⟩ ∈ ker L0 : +P0 |ψ⟩ = |ψ⟩ , +|ψ⟩ ∈ (ker L0)⊥ : +P0 |ψ⟩ = 0. +(8.44) +Then, the appropriate contracting homotopy operator reads ∆(1−P0) and (8.37) is changed +to: +{QB, ∆(1 − P0)} = (1 − P0). +(8.45) +This parallels completely the definition of the Green function in presence of zero-modes, see +(B.3). By abuse of language, we will also say that ∆ is a contracting homotopy operator, +remembering that this statement is correct only when multiplying with (1 − P0). +We will revisit these aspects later from the SFT perspective. In fact, we will find that QB +is the kinetic operator of the gauge invariant theory, while ∆ is the gauge fixed propagator +in the Siegel gauge. This is expected from experience with standard gauge theories: the +inverse of the kinetic operator (Green function) is not defined when the gauge invariance is +not fixed. +The on-shell condition (8.35) is already a good starting point. In order to simplify the +analysis further, one can restrict the question of computing the cohomology on the subspace: +H0 := H ∩ ker b0 = Hm ⊗ Hgh,0, +(8.46) +where Hgh,0 = Hgh ∩ker b0 was defined in (7.2.6). This subspace contains all states |ψ⟩ such +that: +|ψ⟩ ∈ H0 +=⇒ +b0 |ψ⟩ = 0. +(8.47) +In this subspace, there is no exact state |ψ⟩ with L0 |ψ⟩ ̸= 0 such that b0 |ψ⟩ = QB |ψ⟩ = 0. +Indeed, assuming these conditions, (8.42) leads to a contraction: +b0 |ψ⟩ = QB |ψ⟩ = 0, +L0 |ψ⟩ ̸= 0 +=⇒ +|ψ⟩ = 0. +(8.48) +140 + +Note that the converse statement is not true: there are on-shell states such that b0 |ψ⟩ ̸= 0. +This also makes sense because the ghost Hilbert space can be decomposed with respect to +the ghost zero-modes. The cohomology of QB in the subspace H0 is called the relative +cohomology: +Hrel(QB) := H0(QB) = +� +|ψ⟩ ∈ H0 +�� QB |ψ⟩ = 0, ∄ |χ⟩ ∈ H +�� |ψ⟩ = QB |χ⟩ +� +. +(8.49) +The advantage of the subspace b0 = 0 is to precisely pick the representative of Habs which +lies in ker L0. In particular, the operator L0 is simple and has a direct physical interpretation +as the worldsheet Hamiltonian. This condition is also meaningful in string theory because +these states are also mass eigenstates, which have a nice spacetime interpretation, and it +will later be interpreted in SFT as fixing the Siegel gauge. Moreover, it is implied by the +choice of ∆ in (8.39) as the contracting homotopy operator, which is particularly convenient +to work with to derive the cohomology. However, there are other possible choices, which are +interpreted as different gauge fixings. +After having built this cohomology, we can look for the full cohomology by relaxing the +condition b0 = 0. In view of the structure of the ghost Hilbert space (7.169), one can expect +that Habs(QB) = Hrel(QB) ⊕ c0Hrel(QB), which is indeed the correct answer. But, we will +see (building on Section 3.2.2) that, in fact, it is this cohomology which contains the physical +states in string theory, instead of the absolute cohomology. +As a summary, we are looking for QB-closed non-exact states annihilated by b0 and L0: +QB |ψ⟩ = 0, +L0 |ψ⟩ = 0, +b0 |ψ⟩ = 0. +(8.50) +8.3.2 +Relative cohomology +In (8.14a), the BRST operator was decomposed as: +QB = c0L0 − b0M + �QB, +�Q2 +B = L0M. +(8.51) +This shows that, on the subspace L0 = b0 = 0, �QB is nilpotent and equivalent to QB: +|ψ⟩ ∈ H0 ∩ ker L0 +=⇒ +QB |ψ⟩ = �QB |ψ⟩ , +�Q2 +B |ψ⟩ = 0. +(8.52) +Hence, this implies that �QB is a proper BRST operator and the relative cohomology of QB +is isomorphic to the cohomology of �QB: +H0(QB) = H0( �QB). +(8.53) +Next, we introduce light-cone coordinates in the target spacetime. While it does not allow +to write Lorentz covariant expressions, it is helpful mathematically because it introduces a +grading of the Hilbert space, for which powerful theorems exist (even if we will need only +basic facts for our purpose). +Light-cone parametrization +The two scalar fields X0 and X1 are combined in a light-cone (if ϵ0 = −1) or complex (if +ϵ0 = 1) fashion: +X± +L = +1 +√ +2 +� +X0 +L ± +i +√ϵ0 +X1 +L +� +. +(8.54) +141 + +The modes of X± are found by following (7.50):2 +α± +n = +1 +√ +2 +� +α0 +n ± +i +√ϵ0 +α1 +n +� +, +n ̸= 0, +(8.55a) +x± +L = +1 +√ +2 +� +x0 +L ± +i +√ϵ0 +x1 +L +� +, +p± +L = +1 +√ +2 +� +p0 +L ± +i +√ϵ0 +p1 +L +� +, +(8.55b) +The non-zero commutation relations are: +[α+ +m, α− +n ] = ϵ0 m δm+n,0, +[x± +L, p∓ +L] = iϵ0. +(8.56) +This implies that negative-frequency (creation) modes α± +−n are canonically conjugate to +positive-frequency (annihilation) modes α∓ +n . Note the similarity with the first-order system +(7.134). +For later purposes, it is useful to note the following relations: +2 p+ +Lp− +L = (p0 +L)2 + ϵ0(p1 +L)2 = ϵ0 p2 +∥,L, +(8.57a) +x+p− + x−p+ = x0p0 + ϵ0 x1p1, +(8.57b) +� +n +α+ +n α− +m−n = 1 +2 +� +n +� +α0 +nα0 +m−n + ϵ0 α1 +nα1 +m−n +� +. +(8.57c) +In view of the commutators (8.56), the appropriate definitions of the light-cone number +N ± +n and level operators N ± are: +N ± +n = ϵ0 +n α± +−nα∓ +n , +N ± = +� +n>0 +n N ± +n . +(8.58) +The insertion of ϵ0 follows (7.63). Then, one finds the following relation: +N + + N − = N 0 + N 1. +(8.59) +Using these definitions, the variables appearing in L0 (8.33) +L0 = +� +L⊥ +0 − m2 +∥,Lℓ2 − 1 +� ++ �L∥ +0 +(8.60) +can be rewritten as: +m2 +∥,L = −2ϵ0 p+ +Lp− +L, +�L∥ +0 = N + + N − + N b + N c. +(8.61) +The expression for the sum of the Virasoro operators (7.65) easily follows from (8.57): +L0 +m + L1 +m = ϵ0 +� +n +:α+ +n α− +m−n: = ϵ0 +� +n̸=0,m +:α+ +n α− +m−n: + ϵ0 +� +α− +0 α+ +m + α+ +mα− +m +� +. +(8.62) +Computation – Equation (8.56) +For the modes α± +m, we have: +[α+ +m, α± +n ] = 1 +2 +�� +α0 +m + +i +√ϵ0 +α1 +m +� +, +� +α0 +n ± +i +√ϵ0 +α1 +n +�� += 1 +2 +� +[α0 +m, α0 +n] ∓ 1 +ϵ0 +[α1 +m, α1 +n] +� += ϵ0 +2 m δm+n,0(1 ∓ 1), +2For ϵ0 = 1, this convention matches the ones from [29] for X0 = X and X1 = φ. For ϵ = −1, this +convention matches [193]. +142 + +where we used (7.72). The other commutators follow similarly from (7.73), for example: +[x− +L, p± +L] = 1 +2 +�� +x0 +L − +i +√ϵ0 +x1 +L +� +, +� +p0 +L ± +i +√ϵ0 +p1 +L +�� += 1 +2 +� +[x0 +L, p0 +L] ± ϵ0[x1 +L, p1 +L] +� += ϵ0 +2 (1 ± 1). +Computation – Equation (8.57) +For the modes α± +m, we have: +� +n +α+ +n α− +m−n = 1 +2 +� +n +� +α0 +n + +i +√ϵ0 +α1 +n +� � +α0 +m−n − +i +√ϵ0 +α1 +m−n +� += 1 +2 +� +n +� +α0 +nα0 +m−n + ϵ0 α1 +nα1 +m−n + +i +√ϵ0 +(α0 +m−nα1 +n − α0 +nα1 +m−n) +� +. +The last two terms in parenthesis cancel as can be seen by shifting the sum n → m − n +in one of the term. Note that, for m ̸= 2n, there is no cross-term only after summing +over n. +The relations for the zero-modes follow simply by observing that expressions in both +coordinates can be rewritten in terms of the 2-dimensional (spacetime) flat metric. +Computation – Equation (8.59) +Using (8.57), one finds: +N 0 + N 1 = +� +n +n +� +N 0 +n + N 1 +n +� += +� +n +n +� +N + +n + N − +n +� += N + + N −. +Reduced cohomology +In terms of the light-cone variables, the reduced BRST operator �QB reads: +�QB = +� +m̸=0 +c−m +� +L⊥ +m + ϵ0 +� +n +α+ +n α− +m−n +� ++ 1 +2 +� +m,n +(n − m) :c−mc−nbm+n:. +(8.63) +This operator can be further decomposed. Introducing the degree +deg := N + − N − + � +N c − � +N b +(8.64) +such that +∀m ̸= 0 : +deg(α+ +m) = deg(cm) = 1, +deg(α− +m) = deg(bm) = −1, +(8.65) +and deg = 0 for the other variables, the operator �QB is decomposed as:3 +�QB = Q0 + Q1 + Q2, +deg(Qj) = j, +(8.66a) +where +Q1 = +� +m̸=0 +c−mL⊥ +m + +� +m,n̸=0 +m+n̸=0 +⋆ +⋆c−m +� +ϵ0 α+ +n α− +m−n + 1 +2 (m − n) c−mbm+n +� +⋆ +⋆, +Q0 = +� +n̸=0 +α+ +0 c−nα− +n , +Q2 = +� +n̸=0 +α− +0 c−nα+ +n . +(8.66b) +3The general idea behind this decomposition is the notion of filtration, nicely explained in [3, sec. 3, 36]. +143 + +The nilpotency of �QB implies the following conditions on the Qj: +Q2 +0 = Q2 +2 = 0, +{Q0, Q1} = {Q1, Q2} = 0, +Q2 +1 + {Q0, Q2} = 0. +(8.67) +Hence, Q0 and Q2 are both nilpotent and define a cohomology. +One can show that the cohomologies of �QB and Q0 are isomorphic4 +H0( �QB) ≃ H0(Q0) +(8.68) +under general conditions [29], in particular, if the cohomology is ghost-free (i.e. all states +have Ngh = 1). +The contracting homotopy operator for Q0 is +�∆ := B +�L∥ +0 +, +B := ϵ0 +� +n̸=0 +1 +α+ +0 +α+ +−nbn. +(8.69) +Indeed, it is straightforward to check that +�L∥ +0 = {Q0, B} +=⇒ +{Q0, �∆} = 1. +(8.70) +As a consequence, a necessary condition for a closed �L∥ +0-eigenstate |ψ⟩ to be in the +cohomology of Q0 is to be annihilated by �L∥ +0: +�L∥ +0 |ψ⟩ = 0, +=⇒ +N ± |ψ⟩ = N c |ψ⟩ = N b |ψ⟩ = 0, +(8.71) +since �L∥ +0 is a sum of positive integers. This means that the state ψ contains no ghost or +light-cone excitations α± +−n, b−n and c−n, and lies in the ground state of the Fock space H∥,0. +Then, we need to prove that this condition is sufficient: states with �L∥ +0 = 0 are closed. +First, note that a state |ψ⟩ ∈ H0 with �L∥ +0 has ghost number 1 since there are no ghost +excitations on top of the vacuum | ↓⟩, which has Ngh = 1. Second, �L0 and Q0 commute, +such that: +0 = Q0�L∥ +0 |ψ⟩ = �L∥ +0Q0 |ψ⟩ . +(8.72) +Since Q0 increases the ghost number by 1, one can invert �L∥ +0 = N b + N c + · · · in the last +term since �L∥ +0 ̸= 0 in this subspace. This gives: +Q0 |ψ⟩ = 0. +(8.73) +Hence, the condition �L∥ +0 |ψ⟩ = 0 is sufficient for |ψ⟩ to be in the cohomology. This has to be +contrasted with Section 8.3.1 where the condition L∥ +0 = 0 is necessary but not sufficient. +In this case, the on-shell condition (8.33) reduces to +L0 = L⊥ +0 − m2 +∥,Lℓ2 − 1 = 0. +(8.74) +But, additional states can be found in ker B or in a subspace of H on which B is singular. +We have ker B = ker �L∥ +0 such that nothing new can be found there. However, the operator B +is not defined for states with vanishing momentum α+ +0 ∝ p+ +L = 0. In fact, one must also have +α− +0 ∝ p− +L = 0 (otherwise, the contracting operator for Q2 is well-defined and can be used +instead). But, these states do not satisfy the on-shell condition (except for massless states +with L⊥ +0 = 1), as it will be clear later (see [245, sec. 2.2] for more details). For this reason, +we assume that states have a generic non-zero momentum and that there is no pathology. +4The role of Q0 and Q2 can be reversed by changing the sign in the definition of the degree and the role +of P ± +n . +144 + +Full relative cohomology +This section aims to construct states in H0( �QB) from states in H(Q0). +We follow the +construction from [29]. +Given a state |ψ0⟩ ∈ H0(Q0), the state Q1 |ψ0⟩ is Q0-closed since Q0 and Q1 anticommute +(8.67): +{Q0, Q1} |ψ0⟩ = 0 +=⇒ +Q0 +� +Q1 |ψ0⟩ +� += 0. +(8.75) +Since Q1 |ψ0⟩ is not in ker �L∥ +0 (because Q1 increases the ghost number by 1), the state Q1 |ψ0⟩ +is Q0-exact and can be written as Q0 of another state |ψ1⟩: +Q1 |ψ0⟩ =: −Q0 |ψ1⟩ +=⇒ +|ψ1⟩ = − B +�L∥ +0 +Q1 |ψ0⟩ . +(8.76) +Computation – Equation (8.76) +Start from the definition and insert (8.70) since �L0 is invertible: +Q1 |ψ0⟩ = +� +Q0, B +�L∥ +0 +� +Q1 |ψ0⟩ = Q0 +� +B +�L∥ +0 +Q1 |ψ0⟩ +� +. +The state |ψ1⟩ is identified with minus the state inside the parenthesis (up to a BRST +exact state). +As for |ψ0⟩, apply {Q0, Q1} on ψ1: +{Q0, Q1} |ψ1⟩ = Q0 +� +Q1 |ψ1⟩ + Q2 |ψ0⟩ +� +. +(8.77) +This implies that the combination in parenthesis is Q0-closed and, for the same reason as +above, it is exact: +Q1 |ψ1⟩ + Q2 |ψ0⟩ = Q0 |ψ2⟩ , +|ψ2⟩ = − B +�L∥ +0 +� +Q1 |ψ1⟩ + Q2 |ψ0⟩ +� +. +(8.78) +Computation – Equation (8.77) +{Q0, Q1} |ψ1⟩ = Q0Q1 |ψ1⟩ − Q2 +1 |ψ0⟩ = Q0Q1 |ψ1⟩ + {Q0, Q2} |ψ0⟩ . +The first equality follows from (8.76), the second by using (8.67). The final result is +obtained after using that |ψ0⟩ is Q0-closed. +Iterating this procedure leads to a series of states: +|ψk+1⟩ = − B +�L∥ +0 +� +Q1 |ψk⟩ + Q2 |ψk−1⟩ +� +. +(8.79) +We claim that a state in the relative cohomology |ψ⟩ ∈ H0( �QB) is built by summing all +these states: +|ψ⟩ = +� +k∈N +|ψk⟩ . +(8.80) +Indeed, it is easy to check that |ψ⟩ is �QB-closed: +�QB |ψ⟩ = 0. +(8.81) +145 + +We leave aside the proof that ψ is not exact (see [29]). Note that ψ and ψ0 have the same +ghost numbers +Ngh(ψ) = Ngh(ψ0) = 1 +(8.82) +since Ngh(BQj) = 0. +In fact, since ψ0 does not contain longitudinal modes, it is annihilated by Q1 and Q2 +(these operators contain either a ghost creation operator together with a light-cone annihil- +ation operator, or the reverse): +Q1 |ψ0⟩ = Q2 |ψ0⟩ = 0. +(8.83) +As a consequence, one has ψk = 0 for k ≥ 1 and ψ = ψ0. +Computation – Equation (8.81) +�QB |ψ⟩ = +� +k∈N +�QB |ψk⟩ += Q0 |ψ0⟩ + Q1 |ψ0⟩ + Q0 |ψ1⟩ +� +�� +� +=0 ++ Q2 |ψ0⟩ + Q1 |ψ1⟩ + Q0 |ψ2⟩ +� +�� +� +=0 ++ · · · += 0. +8.3.3 +Absolute cohomology, states and no-ghost theorem +The absolute cohomology is constructed from the relative cohomology: +Habs(QB) = Hrel(QB) ⊕ c0 Hrel(QB). +(8.84) +The interested reader is refereed to [29] for the proof. A simple motivation is that the Hilbert +space is decomposed in terms of the ghost zero-modes as in (7.169). Since the zero-modes +commute with �Q0, linear combination of states in Hrel(QB) and c0Hrel(QB) are expected +to be in the cohomology. Obviously, one has to work out the other terms of QB and prove +that there are no other states. +It looks like there is a doubling of the physical states, one built on | ↓⟩ and one on | ↑⟩. +The remedy is to impose the condition b0 = 0 on the states (see also Section 3.2.2 and [245, +sec. 2.2] for more details). As already pointed out, states in Habs form equivalence class +under |ψ⟩ ∼ |ψ⟩ + QB |Λ⟩, and it is necessary to select a single representative. This is what +the condition b0 = 0 achieves. Obviously, it is always possible to add BRST exact states to +write another representative (for example, to restore the Lorentz covariance). +The last step is to discuss the no-ghost theorem: the latter states that there is no +negative-norm states in the BRST cohomology of string theory. This follows straightfor- +wardly from the condition �L0 = 0: it implies that there are no ghost and no light-cone +excitations. The ghosts and the time direction (if X0 is timelike) are responsible for negative- +norm states. Hence, the cohomology has no negative-norm states if the transverse CFT is +unitary (which implies that all states in H⊥ have a positive-definite inner-product). +Physical states |ψ⟩ ∈ Hrel(QB) are thus of the form: +|ψ⟩ = |k0, k1, ↓⟩ ⊗ |ψ⊥⟩ , +|ψ⊥⟩ ∈ H⊥, +(8.85a) +� +L⊥ +0 − m2 +∥,Lℓ2 − 1 +� +|ψ⟩ = 0, +p2 +L,∥ = −m2 +∥,Lℓ2. +(8.85b) +This form can be made covariant: taking a state of the form |ψ⟩⊗| ↓⟩ with |ψ⟩ ∈ Hm, acting +with QB implies the equivalence with the old covariant quantization: +(Lm +0 − 1) |ψ⟩ = 0, +∀n > 0 : +Lm +n |ψ⟩ = 0. +(8.86) +146 + +This means that ψ must be a weight 1 primary field of the matter CFT. +Remark 8.2 (Open string) The results of this section provide, in fact, the cohomology +for the open string after taking pL = p (instead of pL = p/2 for the closed string). +8.3.4 +Cohomology for holomorphic and anti-holomorphic sectors +It remains to generalize the computation of the cohomology when considering both the +holomorphic and anti-holomorphic sectors. +In this case, the BRST operator is +QB = c0L0 − b0M + �QB + ¯c0 ¯L0 − ¯b0 ¯ +M + �QB. +(8.87) +It is useful to rewrite this expression in terms of L± +0 , b± +0 and c± +0 : +QB = c+ +0 L+ +0 − b+ +0 M + + c− +0 L− +0 − b− +0 M − + �Q+ +B, +(8.88) +where +L+ +0 = +� +L⊥+ +0 +− +m2 +∥ℓ2 +2 +− 2 +� ++ �L∥+ +0 , +L− +0 = L⊥− +0 ++ �L∥− +0 +(8.89) +and +M ± := 1 +2(M ± ¯ +M). +(8.90) +Because of the relations L± +0 = {QB, b± +0 }, we find that states in the cohomology must be +on-shell L+ +0 = 0 and must satisfy the level-matching condition L− +0 = 0:5 +L+ +0 |ψ⟩ = L− +0 |ψ⟩ = 0. +(8.91) +Again, it is possible to reduce the cohomology by imposing conditions on the zero-modes +such that the above conditions are automatically satisfied (see also Section 3.2.2). Imposing +first the condition b− +0 = 0 defines the semi-relative cohomology. The relative cohomology is +found by imposing b± +0 = 0 and in fact corresponds to the physical space (see [245, sec. 2.3] for +more details). The rest of the derivation follows straightforwardly because the two sectors +commute: we find that the cohomology is ghost-free and has no light-cone excitations: +�L∥± +0 += N 0 ± ¯N 0 + N 1 ± ¯N 1 + N b ± ¯N b + N c ± ¯N c = 0. +(8.92) +In general, it is simpler to work with a covariant expression and to impose the necessary +conditions. Taking a state |ψ⟩ ⊗ | ↓↓⟩ with |ψ⟩ ∈ Hm, we find that ψ is a weight (1, 1) +primary field of the matter CFT: +(Lm +0 + ¯Lm +0 − 2) |ψ⟩ = 0, +(Lm +0 − ¯Lm +0 ) |ψ⟩ = 0, +∀n > 0 : +Lm +n |ψ⟩ = ¯Lm +n |ψ⟩ = 0. +(8.93) +An important point is that the usual mass-shell condition k2 = −m2 is provided by the first +condition only. This also shows that states in the cohomology naturally appears with c¯c +insertion since +| ↓↓⟩ = c(0)¯c(0) |0⟩ = c1¯c1 |0⟩ . +(8.94) +This hints at rewriting of scattering amplitudes in terms of unintegrated states (3.29) only. +A state is said to be of level (ℓ, ¯ℓ) and denoted as ψℓ,¯ℓ if it satisfies: +�L0 |ψℓ,¯ℓ⟩ = ℓ |ψℓ,¯ℓ⟩ , +�¯L0 |ψℓ,¯ℓ⟩ = ¯ℓ |ψℓ,¯ℓ⟩ . +(8.95) +5In the current case, the propagator is less easily identified. We will come back on its definition later. +147 + +Example 8.1 – Closed string tachyon +As an example, let’s construct the state ψ0,0 with level zero for a spacetime with D +non-compact dimensions. In this case, the transverse CFT contains D − 2 free scalars +which combine with X0 and X1 into D scalars Xµ. The Fock space is built on the +vacuum |k⟩ and we define the mass such that on-shell condition reduces to the standard +QFT expression: +k2 = −m2, +m2 := 2 +ℓ2 (N + ¯N − 2), +(8.96) +where N and ¯N are the matter level operators. The state in the remaining transverse +CFT (without the D − 2 scalars) is the SL(2, C) vacuum with L⊥ +0 = ¯L⊥ +0 = 0 (this is +the state with the lowest energy for a unitary CFT). In this case, the on-shell condition +reads +m2ℓ2 = −4 < 0. +(8.97) +Since the mass is negative, this state is a tachyon. The vertex operator associated to +the state reads: +V (k, z, ¯z) = c(z)¯c(¯z)eik·X(z,¯z). +(8.98) +8.4 +Summary +In this chapter, we have described the BRST quantization from the CFT point of view. We +have first considered only the holomorphic sector (equivalently, the open string). We proved +that the cohomology does not contain negative-norm states and we provided an explicit way +to construct the states. Finally, we glued together both sectors and characterized the BRST +cohomology of the closed string. +What is the next step? We could move to computations of on-shell string amplitudes, +but this falls outside the scope of this book. We can also start to consider string field theory. +Indeed, the BRST equation QB |ψ⟩ = 0 and the equivalence |ψ⟩ ∼ |ψ⟩ + QB |Λ⟩ completely +characterize the states. In QFT, states are solutions of the linearized equations of motion: +hence, the BRST equation can provide a starting point for building the action. This is the +topic of Chapter 10. +8.5 +Suggested readings +• The general method to construct the absolute cohomology follows [29, 193]. Other +works and reviews include [20, 28, 57, 118, 119, 170, 177]. +• String states are discussed in [24, sec. 3.3, 193, sec. 4.1]. +148 + +Part II +String field theory +149 + +Chapter 9 +String field +In this chapter, we introduce general concepts about the string field. The goal is to give an +idea of which type of object it is and of the different possibilities for describing it. We will +see that the string field is a functional and, for this reason, it is more convenient to work +with the associated ket field, which can itself be represented in momentum space. We focus +on what to expect from a free field, taking inspiration from the worldsheet theory. The +interpretation becomes more difficult when taking into account the interactions. +9.1 +Field functional +A string field, after quantization, is an operator which creates or destroys a string at a given +time. Since a string is a 1-dimension extended object, the string field Ψ must depend on the +spatial positions of each point of the string denoted collectively as Xµ. Hence, the string +field is a functional Ψ[Xµ]. The fact that it is a functional rather than a function makes the +construction of a field theory much more challenging: it asks for revisiting all concepts we +know in point-particle QFT without any prior experience with a simple model.1 +It is important that the dependence is only on the shape and not on the parametrization. +However, it is simpler to first work with a specific parametrization X(σ) and make sure that +nothing depends on it at the end (equivalent to imposing the invariance under reparamet- +rization of the worldsheet). This leads to work with a functional Ψ[X(σ)] of fields on the +worldsheet (at fixed time). To proceed, one should first determine the degrees of freedom of +the string, and then to find the interactions. The simplest way to achieve the first step is to +perform a second-quantization of the string wave-functional: the string field is written as a +linear combination of first-quantized states with spacetime wave functionals as coefficients.2 +This provides a free Hamiltonian; trying to add interactions perturbatively does not work +well. +It is not possible to go very far with this approach and one is lead to choose a specific +gauge, breaking the manifest invariance under reparametrizations. The simplest is the light- +cone gauge since one works only with the physical degrees of freedom of the string. While +this approach is interesting to gain some intuitions and to show that, in principle, it is +possible to build a string field theory, it requires making various assumptions and ends up +1The problem is not in working with the wordline formalism and writing a BRST field theory, but really +to take into account the spatial extension of the objects. +In fact, generalizing further to functionals of +extended (p > 1)-dimensional objects – branes – shows that SFT is the simplest of such field theories. +2The description of the first-quantized states depends on the CFT used to describe the theory. This +explains the lack of manifest background independence of SFT. Unfortunately, no better approach has been +found until now. +150 + +with problems (especially for superstrings).3 +Since worldsheet reparametrization invariance is just a kind of gauge symmetry – maybe +less familiar than the non-Abelian gauge symmetries in Yang–Mills, but still a gauge sym- +metry –, one may surmise that it should be possible to gauge fix this symmetry and to +introduce a BRST symmetry in its place. This is the program of the BRST (or covariant) +string field theory in which the string field depends not only of the worldsheet (at fixed +time), but also on the ghosts: Ψ[X(σ), c(σ)]. There is no dependence on the b ghost because +the latter is the conjugate momentum of the c ghost: in the operator language, b(σ) ∼ +δ +δc(σ). +The BRST formalism has the major advantage to allow to move easily from D = 26 +dimensions – described by Xµ scalars (µ = 0, . . . , 25) – to a (possibly curved) D-dimensional +spacetime and a string with some internal structure – described by a more general CFT, +in which D scalars Xµ represent the non-compact dimensions and the remaining system +with central charge 26 − D describes the compactification and structure. It is sufficient to +consider the string field as a general functional of all the worldsheet fields. For simplicity, +we will continue to write X in the functional dependence, keeping the other matter fields +implicit. +It is complicated to find an explicit expression for the string field as a functional of X(σ) +and c(σ). In fact, the field written in this way is in the position representation and, as usual +in quantum mechanics, one can choose to work with the representation independent ket |Ψ⟩: +Ψ[X(σ), c(σ)] := ⟨X(σ), c(σ)|Ψ⟩ . +(9.1) +It is often more convenient to work with |Ψ⟩ (which we will also denote simply as Ψ, not +distinguishing between states and operators). The latter will be the basic object of SFT in +most of this book. +Writing a field theory in terms of |Ψ⟩ may not be intuitive since in point-particle QFT, +one is used to work with the position or momentum representation. In fact, there is a very +simple way to recover a formulation in terms of spacetime point-particle fields, which can be +used almost whenever there is a doubt about what is going on. Indeed, as is well-known from +standard worldsheet string theory, the string states behave like a collection of particles. This +is because the modes of the CFT fields (like αµ +n) carry spacetime indices (Lorentz, group +representation. . . ) such that the states themselves carries indices. Indeed, these quantum +numbers classify eigenstates of the operators L0 and ¯L0. On the other hand, positions and +shapes are not eigenstates of any simple CFT operator. +9.2 +Field expansion +It follows that the second-quantized string field can be written as a linear combination +of first-quantized off-shell states |φα(k)⟩ = Vα(k; 0, 0) |0⟩ (which form a basis of the CFT +Hilbert space H): +|Ψ⟩ = +� +α +� +dDk +(2π)D ψα(k) |φα(k)⟩ , +(9.2) +where k is the D-dimensional momentum of the string (conjugated to the position of the +centre-of-mass) and α is a collection of discrete quantum numbers (Lorentz indices, group +representation. . . ). When inserting this expansion inside the action, we find that it reduces +to a standard field theory with an infinite number of particles described by the spacetime +3While this approach has been mostly abandoned, recent results show that it can still be used when +defined with a proper regularization [4, 114–117]. +151 + +fields ψα(k) (momentum representation). +The fields can also be written in the position +representation by Fourier transforming only the momentum k to the centre-of-mass x: +ψα(x) = +� +dDk +(2π)D eik·xψα(k). +(9.3) +However, we will see that it is often not convenient because the action is non-local in position +space (including for example exponentials of derivatives). +The physical intuition is that the string is a non-local object in spacetime. It can be +expressed in momentum space through a Fourier transformation: variables dual to non- +compact (resp. compact) dimensions are continuous (discrete). As a consequence, the mo- +mentum is continuous since the centre-of-mass move in the non-compact spacetime, while +the string itself has a finite extension and the associated modes are discrete but still not +bounded (and similarly for compact dimensions). This indicates that the spectrum is the +collection of a set of continuous and discrete modes. Hence, the non-locality of the string +(due to the spatial extension) is traded for an infinite number of modes which behave like +standard particles. In this description, the non-locality arises: 1) in the infinite number of +fields, 2) in the coupling between the modes, 3) as a complicated momentum-dependence of +the action. +When we are not interested in the spacetime properties, we will write a generic basis of +the Hilbert space H as {φr}: +|Ψ⟩ = +� +r +ψr |φr⟩ . +(9.4) +The sum over r includes discrete and continuous labels. +Example 9.1 – Scalar field +In order to illustrate the notations for a point-particle, consider a scalar field φ(x). It +can be expanded in Fourier modes as: +φ(x) = +� +dDk +(2π)D φ(k)eik·x. +(9.5) +The corresponding ket |φ⟩ is found by expanding on a basis {|k⟩}: +|φ⟩ = +� +dDk +(2π)D φ(k) |k⟩ , +φ(k) = ⟨k|φ⟩ . +(9.6) +Similarly, the position space field is defined from the basis {|x⟩} such that: +φ(x) = ⟨x|φ⟩ = +� +dDk +(2π)D ⟨x|k⟩ ⟨k|φ⟩ , +⟨x|k⟩ = eik·x. +(9.7) +9.3 +Summary +In this chapter, we introduced general ideas about what a string field is. We now need to +write an action. In general, one proceeds in two steps: +1. build the kinetic term (free theory): +(a) equations of motion → physical states +(b) equivalence relation → gauge symmetry +2. add interactions and deform the gauge transformation +152 + +We consider the first point in the next chapter, but we will have to introduce more machinery +in order to discuss interactions. +9.4 +Suggested readings +• General discussions of the string field and of the ideas of string field theory can be +found in [192, sec. 4, 261]. +• Light-cone SFT is reviewed in [245, 124, chap. 6, 125, chap. 9]. +153 + +Chapter 10 +Free BRST string field theory +Abstract +In this chapter, we construct the BRST (or covariant) free bosonic string field +theories. It is useful to first ignore the interactions in order to introduce some general tools +and structures in a simpler setting. Moreover, the free SFT is easily constructed and does +not require as much input as the interactions. In this chapter, we discuss mostly the open +string, keeping the closed string for the last section. We start by describing the classical +theory: equations of motion, action, gauge invariance and gauge fixing. Then, we perform +the path integral quantization and compute the action in terms of spacetime fields for the +first two levels (tachyon and gauge field). +10.1 +Classical action for the open string +Contrary to most of this book, we will exemplify the discussion with the open string. The +reason is that most computations are the same in both the open and closed string theories, +but the latter requires twice more writing. There are also a few subtleties which can be more +easily explained once the general structure is understood. Everything needed for the open +string for this chapter can be found in Chapter 8: in fact, describing the open string (at this +level) is equivalent to consider only the holomorphic sector of the CFT and to set pL = p +(instead of p/2). We consider a generic matter CFT in addition to the ghost system and we +denote as H the space of states. The open and closed string fields are denoted respectively +by Φ and Ψ, such that it is clear which theory is studied. +An action can be either constructed from first principles, or it can be derived from the +equations of motion. Since the fundamental structure of string field theory is not (really) +known, one needs to rely on the second approach. +But do we already know the (free) +equations of motion for the string field? The answer is yes. But, before showing how these +can be found from the worldsheet formalism, we will study the case of the point-particle to +fix ideas and notations. +10.1.1 +Warm-up: point-particle +The free (or linearized) equation of motion for a scalar particle reads: +(−∆ + m2)φ(x) = 0. +(10.1) +Solutions to this equation provides one-particle state of the free theory: a convenient basis +is {eikx}, where each state satisfies the on-shell condition +k2 = −m2. +(10.2) +154 + +The field φ(x) is decomposed on the basis as +φ(x) = +� +dk φ(k)eikx, +(10.3) +where φ(k) are the coefficients of the expansion. Since the field is off-shell, the condition +k2 = −m2 is not imposed. Following Chapter 9, the field can also be represented as a ket: +φ(x) = ⟨x|φ⟩ , +φ(k) = ⟨k|φ⟩ , +(10.4) +or, conversely: +|φ⟩ = +� +dx φ(x) |x⟩ = +� +dk φ(k) |k⟩ . +(10.5) +Writing the kinetic operator as a kernel: +K(x, x′) :=⟨x| K |x′⟩ = δ(x − x′) (−∆x + m2), +(10.6) +the equations of motion reads +� +dx′ K(x, x′)φ(x′) = 0 +⇐⇒ +K |φ⟩ = 0. +(10.7) +An action can easily be found from the equation of motion by multiplying with φ(x) and +integrating: +S = 1 +2 +� +dx φ(x)(−∆ + m2)φ(x) = 1 +2 +� +dxdx′ φ(x)K(x, x′)φ(x′). +(10.8) +It is straightforward to write the action in terms of the ket: +S = 1 +2 ⟨φ| K |φ⟩ . +(10.9) +There is one hidden assumption in the previous lines: the definition of a scalar product. +A natural inner product is provided in the usual quantum mechanics by associating a bra to +a ket. Similarly, integration provides another definition of the inner product when working +with functions. We will find that the definition of the inner product requires more care in +closed SFT. To summarize, to write the kinetic term of the action, one needs the linearized +equation of motion and an appropriate inner product on the space of states. +10.1.2 +Open string action +The worldsheet equation which yields precisely all the string physical states |ψ⟩ is the BRST +condition: +QB |ψ⟩ = 0. +(10.10) +Considering the open string field Φ to be a linear combination of all possible one-string +states |ψ⟩ +Φ ∈ H, +(10.11) +the equation of motion is: +QB |Φ⟩ = 0. +(10.12) +Moving away from the physical state condition, the string field Φ is off-shell and is expanded +on a general basis {φr} of H. This presents a first difficulty because the worldsheet approach +– and the description of amplitudes – looks ill-defined for off-shell states: extending the usual +155 + +formalism will be the topic of Chapter 11. However, this is not necessary for the free theory +and we can directly proceed. +Next, we need to find an inner product ⟨·, ·⟩ on the Hilbert space H. A natural candidate +is the BPZ inner product since it is not degenerate +⟨A, B⟩ := ⟨A|B⟩ , +(10.13) +where ⟨A| = |A⟩t is the BPZ conjugate (6.98) of |A⟩, using I−. This leads to the action: +S = 1 +2 ⟨Φ, QBΦ⟩ = 1 +2 ⟨Φ| QB |Φ⟩ . +(10.14) +Due to the definition of the BPZ product, the action is equivalent to a 2-point correlation +function on the disk. +The inner product satisfies the following identities: +⟨A, B⟩ = (−1)|A||B|⟨B, A⟩, +⟨QBA, B⟩ = −(−1)|A|⟨A, QBB⟩, +(10.15) +where |A| denotes the Grassmann parity of the operator A. +A first consistency check is to verify that the ghost number of the string can be defined +such that the action is not vanishing. Indeed, the ghost number anomaly on the disk implies +that the total ghost number must be Ngh = 3. Since physical states have Ngh = 1, it is +reasonable to take the string field to satisfy the same condition, even off-shell: +Ngh(Φ) = 1. +(10.16) +This condition means that there is no ghost at the classical level beyond the one of the +energy vacuum | ↓⟩, which has Ngh = 1. Moreover, the BRST charge has Ngh(QB) = 1, +such that the action has ghost number 3. +One needs to find the Grassmann parity of the string field. Using the properties of the +BPZ inner product, the string field should be Grassmann odd +|Φ| = 1 +(10.17) +for the action to be even. This is in agreement with the fact that the string field has ghost +number 1 and that the ghosts are Grassmann odd. One must impose a reality condition on +the string field (a complex field would behave like two real fields and have too many states). +The appropriate reality condition identifies the Euclidean and BPZ conjugates: +|Φ⟩‡ = |Φ⟩t . +(10.18) +That this relation is correct will be checked a posteriori for the tachyon field in Section 10.4. +Computation – Equation (10.17) +⟨Φ, QBΦ⟩ = (−1)|Φ|(|QBΦ|)⟨QBΦ, Φ⟩ = (−1)|Φ|(1+|Φ|)⟨QBΦ, Φ⟩ += ⟨QBΦ, Φ⟩ = −(−1)|Φ|⟨Φ, QBΦ⟩, +where both properties (10.15), together with the fact that |Φ|(1 + |Φ|) is necessarily +even. In order for the bracket to be non-zero, one must have |Φ| = 1. +Since the Hilbert space splits as H = H0 ⊕ c0H0 with H0 = H ∩ ker b0, see (8.31), it is +natural to split the field as (this is discussed further in Section 10.2): +|Φ⟩ = |Φ↓⟩ + c0 |�Φ↓⟩ , +(10.19) +156 + +where +Φ↓, �Φ↓ ∈ H0 +=⇒ +b0 |Φ↓⟩ = b0 |�Φ↓⟩ = 0. +(10.20) +The ghost number of each component is +Ngh(Φ↓) = 1, +Ngh(�Φ↓) = 0. +(10.21) +Remembering the decomposition (8.14a) of the BRST operator +QB = c0L0 − b0M + �QB, +(10.22) +inserting the decomposition (10.19) in the action (10.14) gives: +S = 1 +2⟨Φ↓| c0L0 |Φ↓⟩ + 1 +2⟨�Φ↓| c0M |�Φ↓⟩ +⟨�Φ↓| c0 �QB |Φ↓⟩ . +(10.23) +The equations of motion are obtained by varying the different fields: +0 = −M |�Φ↓⟩ + �QB |Φ↓⟩ , +0 = c0L0 |Φ↓⟩ + c0 �QB |�Φ↓⟩ . +(10.24) +Computation – Equation (10.23) +Let’s introduce the projector Πs = b0c0 on the space H0 = H∩ker b0 and the orthogonal +projector ¯Πs = c0b0 such that +|Φ⟩ = |Φ↓⟩ + |Φ↑⟩ , +|Φ↓⟩ = Πs |Φ⟩ , +|Φ↑⟩ = ¯Πs |Φ⟩ . +(10.25) +We then have: +ΠsQB |Φ⟩ = −b0M |Φ↑⟩ + �QB |Φ↓⟩ , +¯ΠsQB |Φ⟩ = c0L0 |Φ↓⟩ + �QB |Φ↑⟩ , +(10.26) +using +[Πs, �QB] = [Πs, M] = [Πs, L0] = 0. +(10.27) +Then, we need the fact that Π† +s = ¯Πs. to compute the action: +S = 1 +2 ⟨Φ, QBΦ⟩ += 1 +2 ⟨ΠsΦ + ¯ΠsΦ, QBΦ⟩ += 1 +2 ⟨ΠsΦ, ¯ΠsQBΦ⟩ + 1 +2 ⟨¯ΠsΦ, ΠsQBΦ⟩ += 1 +2 ⟨Φ↓, c0L0Φ↓ + �QBΦ↑⟩ + 1 +2 ⟨Φ↑, −b0MΦ↑ + �QBΦ↓⟩ += 1 +2 ⟨Φ↓, c0L0Φ↓⟩ + 1 +2 ⟨Φ↓, �QBΦ↑⟩ − 1 +2 ⟨Φ↑, b0MΦ↑⟩ + 1 +2 ⟨Φ↑, �QBΦ↓⟩. +The result follows by setting |Φ↑⟩ = c0 |�Φ⟩, using (10.15) and that the BPZ conjugate +of c0 is −c0. +10.1.3 +Gauge invariance +In writing the action, only the condition that the states are BRST closed has been used. One +needs to interpret the condition that the state are not BRST-exact, or phrased differently +that two states differing by a BRST exact state are equivalent: +|φ⟩ ∼ |ψ⟩ + QB |λ⟩ . +(10.28) +157 + +Uplifting this condition to the string field, the most direct interpretation is that it corres- +ponds to a gauge invariance: +|Φ⟩ −→ |Φ′⟩ = |Φ⟩ + δΛ |Φ⟩ , +δΛ |Φ⟩ = QB |Λ⟩ +Ngh(Λ) = 0. +(10.29) +In order for the ghost numbers to match, the gauge parameter has vanishing ghost number. +The action (10.14) is obviously invariant since the BRST charge is nilpotent. +10.1.4 +Siegel gauge +In writing the action (10.14), the condition b0 |ψ⟩ = 0 has not been imposed on the string +field. In Section 3.2.2, this condition was found by restricting the BRST cohomology, pro- +jecting out states built on the ghost vacuum | ↑⟩, as required by the behaviour of the on-shell +scattering amplitudes. In Chapter 8, we obtained it by finding that the absolute cohomology +contains twice more states as necessary. This was also understood as a way to work with +a specific representative of the BRST cohomology. Since the field is off-shell and since the +action computes off-shell Green functions, these arguments cannot be used, which explains +why we did not use this condition earlier. +On the other hand, the condition +b0 |Φ⟩ = 0 +(10.30) +can be interpreted as a gauge fixing condition, called Siegel gauge. It can be reached from +any field through a gauge transformation (10.29) with +|Λ⟩ = −∆ |Φ⟩ , +∆ = b0 +L0 +, +(10.31) +where ∆ was defined in (8.39) and will be identified with the propagator. Note that b0 = 0 +does not imply L0 = 0 since the string field is not BRST closed. +This gauge choice is well-defined and completely fixes the gauge symmetry off-shell, +meaning that no solution of the equation of motion is pure gauge after the gauge fixing. +This is shown as follows: assume that |ψ⟩ = QB |χ⟩ is an off-shell pure-gauge state with +L0 ̸= 0, then, because it is also annihilated by b0, one finds: +0 = {QB, b0} |ψ⟩ = L0 |ψ⟩ +(10.32) +which yields a contradiction. +The gauge fixing condition breaks down for L0 = 0, but this does not pose any problem +when working with Feynman diagrams since they are not physical by themselves (nor are +the off-shell and on-shell Green functions). Only the sum giving the scattering amplitudes +(truncated on-shell Green functions) is physical: in this case, the singularity L0 = 0 cor- +responds to the on-shell condition and it is well-known how such infrared divergences for +intermediate states are removed (through the LSZ prescription, mass renormalization and +tadpole cancellation). +Computation – Equation (10.31) +Performing a gauge transformation gives +b0 |Φ′⟩ = b0 |Φ⟩ + b0QB |Λ⟩ = 0. +(10.33) +Then, one writes +b0 |Φ⟩ = b0{QB, ∆} |Φ⟩ = b0QB∆ |Φ⟩ , +(10.34) +using the relation (8.37), the expression (8.39) and the fact that b2 +0 = 0. Plugging this +158 + +back in the first equation gives: +b0QB (∆ |Φ⟩ + |Λ⟩) = 0. +(10.35) +The factor of b0 can be removed by multiplying with c0, and the parenthesis should +vanish (since it is not identically closed), which means that (10.31) holds up to a BRST +exact state. +Example 10.1 – Gauge fixing and singularity +In Maxwell theory, the gauge transformation +A′ +µ = Aµ + ∂µλ, +(10.36) +is used to impose the Lorentz condition +∂µA′ +µ = 0 +=⇒ +∆λ = −∂µAµ. +(10.37) +In momentum space, the parameter reads +λ = −kµ +k2 Aµ. +(10.38) +It is singular when k is on-shell, k2 = 0. However, this does not prevent from computing +Feynman diagrams. +To understand the effect of the gauge fixing on the string field components, decompose +the field as (10.19) |Φ⟩ = |Φ↓⟩ + c0 |�Φ↓⟩. Then, imposing the condition (10.30) yields +|�Φ↓⟩ = 0 +=⇒ +|Φ⟩ = |Φ↓⟩ . +(10.39) +This has the expected effect of dividing by two the number of states and show that they are +not physical. +Plugging this condition in the action (10.23) leads to gauge fixed action +S = 1 +2 ⟨Φ| c0L0 |Φ⟩ , +(10.40) +for which the equation of motion is +L0 |Φ⟩ = 0. +(10.41) +But, note that this equation contains much less information than the original (10.12): as |�Φ↓⟩ +is truncated from (10.40), a part of the equations of motion is lost. The missing equation +can be found by setting |�Φ⟩ = 0 in (10.24) and must be imposed on top of the action: +�QB |Φ⟩ = 0. +(10.42) +It is called out-of-Siegel gauge constraint and is equivalent to the Gauss constraint in elec- +tromagnetism: the equations of motion for pure gauge states contain also the physical fields, +thus, when one fixes a gauge, these relations are lost and must be imposed on the side of the +action. This procedure mimics what happens in the old covariant theory, where the Virasoro +constraints are imposed after choosing the flat gauge (if Φ contains no ghost on top of | ↓⟩, +then �QB = 0 implies Ln = 0, see Section 8.3.3). Moreover, the states which do not satisfy +the condition b0 = 0 do not propagate: this restricts the external states to be considered in +amplitudes. +159 + +Remark 10.1 Another way to derive (10.40) is to insert {b0, c0} = 1 in the action: +S = 1 +2 ⟨Φ| QB{c0, b0} |Φ⟩ = 1 +2 ⟨Φ| QBb0c0 |Φ⟩ += 1 +2 ⟨Φ| {b0, QB}c0 |Φ⟩ − 1 +2 ⟨Φ| b0QBc0 |Φ⟩ += 1 +2 ⟨Φ| c0L0 |Φ⟩ . +The drawback of this computation is that it does not show directly how the constraints (10.42) +arise. +Remark 10.2 (Generalized gauge fixing) It is possible to generalize the Siegel gauge, +in the same way that the Feynman gauge generalizes the Lorentz gauge. This has been studied +in [1, 2]. +In this section, we have motivated different properties and adopted some normalizations. +The simplest way to check that they are consistent is to derive the action in terms of the +spacetime fields and to check that it has the expected properties from standard QFT. This +will be the topic of Section 10.4. +10.2 +Open string field expansion, parity and ghost num- +ber +A basis for the off-shell Hilbert space H is denoted by {φr}, where the ghost numbers and +parity of the states are written as: +nr := Ngh(φr), +|φr| = nr +mod 2. +(10.43) +The corresponding basis of dual (or conjugate) states {φc +r} is defined by (6.145): +⟨φc +r|φs⟩ = δrs. +(10.44) +The basis states can be decomposed according to the ghost zero-modes +|φr⟩ = |φ↓,r⟩ + |φ↑,r⟩ , +b0 |φ↓,r⟩ = c0 |φ↑,r⟩ = 0. +(10.45) +Finally, each state ψ↑ ∈ c0H can be associated to a state �ψ: +|ψ↑⟩ = c0 | �ψ↓⟩ , +b0 | �ψ↓⟩ = 0, +Ngh(ψ↑) = Ngh( �ψ↓) + 1. +(10.46) +More details can be found in Section 11.2. +Any field Φ can be expanded as +|Φ⟩ = +� +r +ψr |φr⟩ , +(10.47) +where the ψr are spacetime fields (remembering that r denotes collectively the continuous +and discrete quantum numbers).1 +1The notation is slightly ambiguous: from (10.45), it looks like both components of φr have the same +coefficient ψr. But, in fact, one sums over all linearly independent states: in terms of the components of φr, +different basis can be considered; for example {φ↓,r, φ↑,r}, or {φ↓,r ± φ↑,r}. A more precise expression can +be found in (10.54) and (10.56). +160 + +Obviously, the coefficients do not carry a ghost number since they are not worldsheet +operators. However, they can be Grassmann even or odd such that each term of the sum +has the same parity, so that the field has a definite parity: +∀r : +|Φ| = |ψr| |φr|. +(10.48) +If the field is Grassmann odd (resp. even) then the coefficients ψr and the basis states must +have opposite (resp. identical) parities, such that |Φ| = 1. +Since the parity results from worldsheet ghosts and since there would be Grassmann odd +states even in a purely bosonic theory, it suggests that the parity of the coefficients ψr is +also related to a spacetime ghost number G defined as: +G(ψr) = 1 − nr. +(10.49) +The normalization is chosen such that the component of a classical string field (Ngh = 1) +are classical spacetime fields with G = 0 (no ghost). We will see later that this definition +makes sense. +A quantum string field Φ generally contains components Φn of all worldsheet ghost +numbers n: +Φ = +� +n∈Z +Φn, +Ngh(Φn) = n. +(10.50) +The projections on the positive and negative (cylinder) ghost numbers are denoted by Φ±: +Φ = Φ+ + Φ−, +Φ+ = +� +n>1 +Φn, +Φ− = +� +n≤1 +Φn. +(10.51) +The shift in the indices is explained by the relation (B.56) between the cylinder and plane +ghost numbers. +For a field Φn of fixed ghost number, coefficients of the expansion vanish whenever the +ghost number of the basis state does not match the one of the field: +∀nr ̸= n : +ψr = 0. +(10.52) +Another possibility to define the field Φn is to insert a delta function: +|Φn⟩ = δ(Ngh − n) |Ψ⟩ = +� +r +δ(nr − n) ψr |φr⟩ . +(10.53) +According to (10.45), a string field Φ can also be separated in terms of the ghost zero- +modes: +|Φ⟩ = |Φ↓⟩ + |Φ↑⟩ = |Φ↓⟩ + c0 |�Φ↓⟩ , +(10.54a) +|Φ↑⟩ = c0 |�Φ↓⟩ , +|�Φ↓⟩ = b0 |Φ↑⟩ , +(10.54b) +where the components satisfy the constraints +b0 |Φ↓⟩ = 0, +c0 |Φ↑⟩ = 0, +b0 |�Φ↓⟩ = 0. +(10.55) +The fields |Φ↓⟩ and |Φ↑⟩ (or |�Φ↓⟩) are called the down and top components and they can be +expanded as: +|Φ↓⟩ = +� +r +ψ↓,r |φ↓,r⟩ , +|Φ↑⟩ = +� +r +ψ↑,r |φ↑,r⟩ . +(10.56) +161 + +10.3 +Path integral quantization +The string field theory can be quantized with a path integral: +Z = +� +dΦcl e−S[Φcl] = +� +dΦcl e− 1 +2⟨Φcl|QB|Φcl⟩. +(10.57) +An index has been added to the field to emphasize that it is the classical field (no spacetime +ghosts). The simplest way to define the measure is to use the expansion (9.4) such that +Z = +� � +s +dψs e−S[{ψr}]. +(10.58) +10.3.1 +Tentative Faddeev–Popov gauge fixing +The action can be gauge fixed using the Faddeev–Popov formalism. The gauge fixing con- +dition is +F(Φcl) := b0 |Φcl⟩ = 0. +(10.59) +Its variation under a gauge transformation (10.29) reads +δF = b0QB |Λcl⟩ , +(10.60) +which implies that the Faddeev–Popov determinant is +det δF +δΛcl += det b0QB. +(10.61) +This determinant is rewritten as a path integral by introducing a ghost C and an antighost +B′ string fields (the prime on B′ will become clear below): +det b0QB = +� +dB′dC e−SFP, +SFP = −⟨B′| b0QB |C⟩ . +(10.62) +The ghost numbers are attributed by selecting the same ghost number for the C ghost and for +the gauge parameter, and then requiring that the Faddeev–Popov action is non-vanishing: +Ngh(B′) = 3, +Ngh(C) = 0. +(10.63) +The ghosts can be expanded as +|B′⟩ = δ(Ngh − 3) +� +r +b′ +r |φr⟩ , +|C⟩ = δ(Ngh) +� +r +cr |φr⟩ , +(10.64) +where the coefficients br and cr are Grassmann odd in order for the determinant formula to +make sense: +|br| = |cr| = 1. +(10.65) +Then, since the basis states appearing in B′ and C are respectively odd and even, this +implies +|B′| = 0, +|C| = 1. +(10.66) +However, there is a redundancy in the gauge fixing because the Faddeev–Popov action +is itself invariant under two independent transformations: +δ |C⟩ = QB |Λ−1⟩ , +Ngh(Λ−1) = −1, +(10.67a) +δ |B′⟩ = b0 |Λ′⟩ , +Ngh(Λ′) = 4. +(10.67b) +162 + +This residual invariance arises because not all |Λcl⟩ generate a gauge transformation. Indeed, +if +|Λ⟩ = |Λ0⟩ + QB |Λ−1⟩ , +(10.68) +the field transforms as +|Φ′ +cl⟩ −→ |Φcl⟩ + QB |Λ0⟩ +(10.69) +and there is no trace left of |Λ−1⟩, so it should not be counted. +The second invariance (10.67b) is not problematic because b0 is an algebraic operator (the +Faddeev–Popov action associated to the determinant has no dynamics). The decompositions +of the gauge parameter Λ′ and the B′ field into components (10.54) read: +|B′⟩ = |B′ +↓⟩ + c0 |B⟩ , +|B⟩ := | �B′ +↓⟩ , +(10.70a) +|Λ′⟩ = |Λ′ +↓⟩ + c0 |�Λ′ +↓⟩ . +(10.70b) +The gauge transformations act on the components as: +δ |B′ +↓⟩ = |�Λ′ +↓⟩ , +δ |B⟩ = 0. +(10.71) +This shows that B is gauge invariant and B′ +↓ can be completely removed by the gauge +transformation. This makes sense because B′ +↓ does not appear in the action (10.62). The +gauge transformation (10.67b) can be used to fix the gauge: +|F ′⟩ = c0 |B′⟩ = 0 +=⇒ +|B′ +↓⟩ = 0. +(10.72) +This fixes completely the gauge invariance since the field B is restricted to satisfy b0 |B⟩ = 0, +and the component form (10.71) of the gauge transformation shows that no transformation +is allowed. Moreover, there is no need to introduce a Faddeev–Popov determinant for this +gauge fixing because the corresponding ghosts would not couple to the other fields (and this +would continue to hold even in the presence of interactions, see Remark 10.4). Indeed, from +the absence of derivatives in the gauge transformation, one finds that the determinant is +constant and thus a ghost-representation is not necessary: +det δF ′ +δΛ′ = det c0b0 = det c0 det b0 = 1 +2 det{b0, c0} = 1 +2. +(10.73) +Then, redefining the measure, the partition function and action reduce to +∆FP = +� +dB dC e−SFP[B,C], +SFP =⟨B| QB |C⟩ . +(10.74) +Note that the field B satisfies +b0 |B⟩ = 0, +Ngh(B) = 2, +|B| = 1. +(10.75) +Since both fields are Grassmann odd, the action can be rewritten in a symmetric way: +SFP = 1 +2 +� +⟨B| QB |C⟩ +⟨C| QB |B⟩ +� +. +(10.76) +Remark 10.3 (Ghost and anti-ghost definitions) The definition of the anti-ghost B +and ghost C is appropriate because the worldsheet and spacetime ghost numbers are related +by a minus sign (and a shift of one unit). In the BV formalism, we will see that the fields +contain the matter and ghost fields, while the antifields contain the anti-ghosts. These two +sets are respectively defined with Ngh ≤ 1 and Ngh > 1. +163 + +The constraint b0 |B⟩ = 0 can be lifted by adding a top component: +|B⟩ = |B↓⟩ + c0 | �B↓⟩ +(10.77) +together with the gauge invariance +δ |B⟩ = QB |Λ1⟩ . +(10.78) +Note the difference with (10.70): while B = �B′ +↓ was the top component of the B′ field, here, +it is defined to be the down component, such that |B↓⟩ = | �B′ +↓⟩. However, for the moment, +we keep B to satisfy b0 |B⟩ = 0. +Remark 10.4 (Decoupling of the ghosts) Since the theory is free the Faddeev–Popov +action (10.74) could be ignored and absorbed in the normalization because it does not couple +to the field. On the other hand, when interactions are included, the gauge transformation +is modified and the ghosts couple to the matter fields. +But this is true only for the C +transformation (10.67a), not for (10.67b). Then it means that ghosts introduced for gauge +fixing (10.67b) will never couple to the matter and other ghosts. +The invariance (10.67a) is a gauge invariance for C and must be treated in the same +way as (10.29). Then, following the Faddeev–Popov procedure, one is lead to introduce new +ghosts for the ghosts. But, the same structure appears again. This leads to a residual gauge +invariance, which has the same form. This process continues recursively and one finds an +infinite tower of ghosts. +10.3.2 +Tower of ghosts +In order to simplify the notations, all the fields are denoted by Φn where n gives the ghost +number: +• Φ1 := Φcl is the original physical field +• Φ0 := C and, more generally, Φn with n < 1 are ghosts +• Φ2 := B and, more generally, Φn with n > 1 are anti-ghosts +The recipe is that each pair of ghost fields (Φn+2, Φ−n) is associated to a gauge parameter +Λ−n−1 with n ≥ 0. It is then natural to gather all the fields in a single field +|Φ⟩ = +� +n +|Φn⟩ +(10.79) +satisfying the gauge fixing constraint: +b0 |Φ⟩ = 0 +=⇒ +b0 |Φn⟩ = 0. +(10.80) +For n ≤ 1, these constraints are gauge fixing conditions for the invariance δ |Φn⟩ = QBΛn. +For n > 1, they arise by considering only the top component of the B field. +Finally, the gauge fixing condition can be incorporated inside the action by using a +Lagrange multiplier β, which is an auxiliary string field containing also components of all +ghost numbers: +|β⟩ = +� +n∈Z +|βn⟩ . +(10.81) +The path integral then reads +Z = +� +dΦdβ e−S[Φ,β], +(10.82) +164 + +where +S[Φ, β] = 1 +2⟨Φ| QB |Φ⟩ +⟨β| b0 |Φ⟩ +(10.83a) += +� +n∈Z +�1 +2⟨Φ2−n| QB |Φn⟩ +⟨β4−n| b0 |Φn⟩ +� +. +(10.83b) +The first term of the action has the same form as the classical action (10.14), but now includes +fields at every ghost number. The complete BV analysis is relegated to the interacting theory. +Removing the auxiliary field β = 0, one finds that the action is invariant under the +extended gauge transformation +δ |Φ⟩ = QB |Λ⟩ , +(10.84) +where the gauge parameter has also components of all ghost numbers: +|Λ⟩ = +� +n∈Z +|Λn⟩ . +(10.85) +10.4 +Spacetime action +In order to make the string field action more concrete, and as emphasized in Chapter 9, +it is useful to expand the string field in spacetime fields and to write the action for the +lowest modes. This also helps to check that the normalization chosen until here correctly +reproduces the standard QFT normalizations. For simplicity we focus on the open bosonic +string in D = 26. +We build the string from the vacuum |k, ↓⟩ (Chapter 8) by acting with the ghost positive- +frequency modes b−n and c−n, the zero-mode c0, and from the scalar oscillators iαµ +−n. +Up to level ℓ = 1, the classical open string field can be expanded as +|Φ⟩ = +1 +√ +α′ +� +dDk +(2π)D +� +T(k) + Aµ(k)αµ +−1 + i +� +α′ +2 B(k)b−1c0 + · · · +� +|k, ↓⟩ +(10.86) +before gauge fixing. The spacetime fields are T(k), Aµ(k) and B(k): their roles will be +interpreted below. The first two terms are part of the |Φ↓⟩ component, while the last term +is part of the |Φ↑⟩ component. +All terms are correctly Grassmann even and they have +vanishing spacetime ghost numbers. The normalizations are chosen in order to retrieve the +canonical normalization in QFT. The factor of i in front of B is needed for the field B to +be real (as can be seen below, this leads to the expected factor ikµ which maps to ∂µ in +position space). +The equation (10.12) leads to the following equations of motion of the spacetime fields: +(α′k2 − 1)T(k) = 0, +k2Aµ(k) + ikµB(k) = 0, +kµAµ(k) + iB(k) = 0. +(10.87) +Moreover, plugging the last equation into the second one gives +k2Aµ(k) − kµk · A(k) = 0. +(10.88) +After Fourier transformation, the equations in position space read: +(α′∆ + 1) T = 0, +B = ∂µAµ, +∆Aµ = ∂µB. +(10.89) +This shows that T(k) is a tachyon with mass m2 = −1/α′ and Aµ(k) is a massless gauge +field. The field B(k) is the Nakanishi–Lautrup auxiliary field: it is completely fixed once +Aµ is known since its equation has no derivative. Siegel gauge imposes B = 0 which shows +that it generalizes the Feynman gauge to the string field. +165 + +Computation – Equation (10.87) +Keeping only the levels 0 and 1 terms in the string field, it is sufficient to truncate the +BRST operator as +QB = c0L0 − b0M + �QB, +M ∼ 2c−1c1, +�QB ∼ c1Lm +−1 + c−1Lm +1 , +Lm +1 ∼ α0 · α1, +Lm +−1 ∼ α0 · α−1. +(10.90) +Acting on the string field gives +QB |Φ⟩ = +1 +√ +α′ +� +dDk +(2π)D +� +T(k)c0L0 |k, ↓⟩ + Aµ(k) +� +c0L0 + ηνρc−1αν +1αρ +0 +� +αµ +−1 |k, ↓⟩ ++ i +� +α′ +2 B(k) +� +− 2b0c−1c1 + ηνρc1αν +−1αρ +0 +� +b−1c0 |k, ↓⟩ +� += +1 +√ +α′ +� +dDk +(2π)D +� +T(k)(α′k2 − 1)c0 |k, ↓⟩ ++ Aµ(k) +� +α′k2c0αµ +−1 + +√ +2α′ηνρηµν kρ c−1 +� +|k, ↓⟩ ++ i +� +α′ +2 B(k) +� +2c−1 + +√ +2α′ηνρkραν +−1c0 +� +|k, ↓⟩ +� += +1 +√ +α′ +� +dDk +(2π)D +� +T(k)(α′k2 − 1)c0 |k, ↓⟩ ++ α′� +Aµ(k)k2 + ikµB(k +� +c0αµ +−1 |k, ↓⟩ ++ +√ +2α′ +� +kµAµ(k) + iB(k) +� +c−1 |k, ↓⟩ +� +. +One needs to be careful when anticommuting the ghosts and we used that pL = k and +α0 = +√ +2α′k for the open string. It remains to require that the coefficient of each state +vanishes. +In order to confirm that Aµ is indeed a gauge field, we must study the gauge transform- +ation. The gauge parameter is expanded at the first level: +|Λ⟩ = +i +√ +2α′ +� +dDk +(2π)D +� +λ(k) b−1 |k, ↓⟩ + · · · +� +. +(10.91) +Note that b−1 | ↓⟩ is the SL(2, C) ghost vacuum. Since +QB |Λ⟩ = +i +√ +α′ +� +dDk +(2π)D λ(k) +� +− +� +α′ +2 k2 b−1c0 + kµαµ +−1 +� +|k, ↓⟩ , +(10.92) +matching the coefficients in (10.29) gives +δAµ = −ikµλ, +δB = k2λ. +(10.93) +This is the appropriate transformation for a U(1) gauge field. +Finally, one can derive the action; for simplicity, we work in the Siegel gauge. We consider +only the tachyon component: +|T⟩ = +� +dDk +(2π)D T(k)c1 |k, 0⟩ , +(10.94) +166 + +with c1 |0⟩ = | ↓⟩. The BPZ conjugate and Hermitian conjugates are respectively: +⟨T| = +� +dDk +(2π)D T(k)⟨−k, 0| c−1, +(10.95a) +⟨T ‡| = +� +dDk +(2π)D T(k)∗⟨k, 0| c−1. +(10.95b) +since ct +1 = c−1 when using the operator I− in (6.111). Imposing equality of both leads to +the reality condition +T(k)∗ = T(−k), +(10.96) +which agrees with the fact that the tachyon is real (the integration measure changes as +dDk → −dDk, but the contour is reversed). +Then, the action reads: +S[T] = 1 +2 +� +dDk +(2π)D T(−k) +� +k2 − 1 +α′ +� +T(k). +(10.97) +This shows that the action is canonically normalized as it should for a real scalar field. +Similarly, one can compute the action for the gauge field: +S[A] = 1 +2 +� +dDk +(2π)D Aµ(−k)k2Aµ(k). +(10.98) +The correct normalization of the tachyon (real scalar field of negative mass) gives a jus- +tification a posteriori for the normalization of the action (10.14). +Typically, string field +actions are normalized in this way, by requiring that the first physical spacetime fields has +the correct normalization. Note how this implies the correct normalization for all the others +physical fields. Generalizing this computation for higher-levels, one always find the kinetic +term to be: +L+ +0 +2 += 1 +2(k2 + m2), +(10.99) +which is the canonical normalization. +Computation – Equation (10.97) +⟨T| c0L0 |T⟩ = 1 +α′ +� +dDk +(2π)D +dDk′ +(2π)D T(k)T(k′)⟨−k′, 0| c−1c0L0c1 |k, 0⟩ += 1 +α′ +� +dDk +(2π)D +dDk′ +(2π)D T(k)T(k′)(α′k2 − 1)⟨−k′, 0| c−1c0c1 |k, 0⟩ += 1 +α′ +� +dDk +(2π)D dDk′ T(k)T(k′)(α′k2 − 1) δ(D)(k + k′), +where we used ⟨0| c−1c0c1 |0⟩ = 1 and ⟨k′|k⟩ = (2π)Dδ(D)(k + k′). +10.5 +Closed string +The derivation of the BRST free action for the closed string is very similar. The starting +point is the equation of motion +QB |Ψ⟩ = 0 +(10.100) +167 + +for the closed string field |Ψ⟩. The difference with (10.12) is that the BRST charge QB now +includes both the left- and right-moving sectors. In the case of the open string, the field Φ +was free of any constraint: we will see shortly that this is not the case for the closed string. +The next step is to find an inner product ⟨·, ·⟩ to write the action: +S = 1 +2 ⟨Ψ, QBΨ⟩. +(10.101) +Following the open string, it seems logical to give the string field Ψ the same ghost number +as the states in the cohomology: +Ngh(Ψ) = 2. +(10.102) +In this case, the ghost number of the arguments of ⟨·, ·⟩ in (10.101) is Ngh = 5. The ghost +number anomaly requires the total ghost number to be 6, that is: +Ngh(⟨·, ·⟩) = 1. +(10.103) +There is no other choice because Ngh(Ψ) must be integer. +The simplest solution is to +insert one c zero-mode c0 or ¯c0, or a linear combination. The BRST operator QB contains +both L± +0 (see the decomposition (8.88)): the natural expectation (and by analogy with the +open string) is that the gauge fixed equation of motion (to be discussed below) should be +equivalent to the on-shell equation L+ +0 = 0 (see also Section 8.3.4). This is possible only if +the insertion is c− +0 . With this insertion, ⟨·, ·⟩ can be formed from the BPZ product: +⟨A, B⟩ =⟨A| c− +0 |B⟩ . +(10.104) +Then, the action reads: +S = 1 +2 ⟨Ψ| c− +0 QB |Ψ⟩ . +(10.105) +However, the presence of c− +0 has a drastic effect because it annihilates part of the string +field. Decomposing the Hilbert space as in (7.175) +H = H− ⊕ c− +0 H−, +H− := H ∩ ker b− +0 , +(10.106) +the string field reads: +|Ψ⟩ = |Ψ−⟩ + c− +0 |�Ψ−⟩ , +Ψ−, �Ψ− ∈ H−, +(10.107) +such that +c− +0 |Ψ⟩ = c− +0 |Ψ−⟩ . +(10.108) +The problem in such cases is that the kinetic term may become non-invertible. This motiv- +ates to project out the component �Ψ− by imposing the following constraint on the string +field: +b− +0 |Ψ⟩ = 0. +(10.109) +The constraint (10.109) is stronger than the constraint L− +0 = 0 for states in the cohomo- +logy (Section 8.3.1), so there is no information lost on-shell by imposing it. For this reason, +we will also impose the level-matching condition: +L− +0 |Ψ⟩ = 0, +(10.110) +such that +Ψ ∈ H− ∩ ker L− +0 . +(10.111) +This will later be motivated by studying the propagator and the off-shell scattering amp- +litudes. To avoid introducing more notations, we will not use a new symbol for this space +and keep implicit that Ψ ∈ ker L− +0 . +168 + +The necessity of this condition can be understood differently. We had found that it is +necessary to ensure that the closed string parametrization is invariant under translations +along the string (Section 3.2.2). Since there is no BRST symmetry associated to this sym- +metry, one needs to keep the constraint.2 This suggests that one may enlarge further the +gauge symmetry and interpret (10.109) as a gauge fixing condition. This would be quite +desirable: one could argue that a fundamental field should be completely described by the +Lagrangian (if such a description exists) and that it should not be necessary to supplement +it with constraints imposed by hand. While this can be achieved at the free level, this idea +runs into problems in the presence of interactions (Section 13.3.1) and the interpretation is +not clear.3 +The action (10.105) is gauge invariant under: +|Ψ⟩ −→ |Ψ′⟩ = |Ψ⟩ + δΛ |Ψ⟩ , +δΛ |Ψ⟩ = QB |Λ⟩ , +(10.112) +where the gauge parameter has ghost number 1 and also lives in H− ∩ ker L− +0 : +Ngh(Λ) = 1, +L− +0 |Λ⟩ = 0, +b− +0 |Λ⟩ = 0. +(10.113) +As for the open string, the gauge invariance (10.112) can be gauge fixed in the Siegel +gauge: +b+ +0 |Ψ⟩ = 0. +(10.114) +Then, the action reduces to: +S = 1 +2 ⟨Ψ| c− +0 c+ +0 L+ +0 |Ψ⟩ = 1 +4 ⟨Ψ| c0¯c0L+ +0 |Ψ⟩ . +(10.115) +The equation of motion is equivalent to the on-shell condition as expected: +L+ +0 |Ψ⟩ = 0. +(10.116) +Additional constraints must be imposed to ensure that only the physical degrees of freedom +propagate. +Computation – Equation (10.115) +c− +0 QB = (c0 − ¯c0)(c0L0 + ¯c0 ¯L0) = c0¯c0(L0 + ¯L0). +10.6 +Summary +In this chapter, we have shown how the BRST conditions defining the cohomology can be +interpreted as an equation of motion for a string field together with a gauge invariance. We +found a subtlety for the closed string due to the ghost number anomaly and because of the +level-matching condition. Then, we studied several basic properties in order to prove that +the free action has the expected properties. +The next step is to add the interactions to the action, but we don’t know first principles +to write them. For this reason, we need to take a detour and to consider off-shell amplitudes. +By introducing a factorization of the amplitudes, it is possible to rewrite them as Feynman +diagrams, where fundamental interactions are connected by propagators (which we will find +to match the one in the Siegel gauge). This can be used to extract the interacting terms of +the action. +2Yet another reason can be found in Section 3.2.2 (see also Section 8.3.4): to motivate the need of the +b+ +0 condition, we could take the on-shell limit from off-shell states because L+ +0 is continuous. However, the +L− +0 operator is discrete and there is no such limit we can consider [245]. So we must always impose this +condition, both off- and on-shell. +3A recent proposal can be found in [179]. +169 + +10.7 +Suggested readings +• The free BRST string field theory is discussed in details in [245] (see also [124, chap. 7, +125, chap. 9, 235, chap. 11]). Shorter discussions can be found in [261, 192, sec. 4, +253, 1, 244]. +• Spacetime fields and actions are discussed in [242, 192, sec. 4]. +• Gauge fixing [147, 242, sec. 6.5, 7.2, 7.4, 1, 134, sec. 2.1, 2, 26]. +• General properties of string field (reality, parity, etc.) [1, 262]. +170 + +Chapter 11 +Introduction to off-shell string +theory +Abstract +In this chapter, we introduce a framework to describe off-shell amplitudes in +string theory. We first start by motivating various concepts – in particular, local coordinates +and factorization – by focusing on the 3- and 4-point amplitudes. We then prepare the stage +for a general description of off-shell amplitudes. We focus again on the closed bosonic string +only. +11.1 +Motivations +11.1.1 +3-point function +The tree-level 3-point amplitude of 3 weight hi vertex operators1 Vi is given by +A0,3 = +� 3 +� +i=1 +Vi(zi) +� +S2 +∝ (z1 − z2)h3−h1−h2 × perms × c.c. +(11.1) +There is no integration since dim M0,3 = 0. +The amplitude is independent of the zi only if the matter state is on-shell, hi = 0, for +example if Vi = c¯cVi with h(Vi) = 1. Indeed, if hi ̸= 0, then A0,3 is not invariant under +conformal transformations (6.38): +z −→ fg(z) = az + b +cz + d ∈ SL(2, C) +(11.2) +(it transforms covariantly). This is a consequence of the punctures: the presence of the latter +modifies locally the metric, since they act as sources of negative curvature. When performing +a conformal transformation, the metric around the punctures changes in a different way as +away from them. This implies that the final result depends on the metric chosen around +the punctures. This looks puzzling because the original path integral derivation (Chapter 3) +indicates that the 3-point amplitude should not depend on the locations of the operators +because its moduli space is empty (hence, all choices of zi should be equivalent). +1The quantum number (k, j) of the vertex operator is mostly irrelevant for the discussion of the current +and next chapters, and they are omitted. +We will distinguish them by a number and reintroduce the +momentum k when necessary. We also omit the overall normalization of the amplitudes. +171 + +The solution is to introduce local coordinates wi with a flat metric |dwi|2 around each +puncture conventionally located at wi = 0. The local coordinates are defined by the maps: +z = fi(wi), +zi = fi(0). +(11.3) +This is also useful to characterize in a simpler way the dependence of off-shell amplitudes +rather than using the metric around the punctures (computations may be more difficult with +a general metric). +The expression of a local operator in the local coordinate system is found by applying +the corresponding change of coordinates (6.48): +f ◦ V (w) = f ′(w)hf ′(w) +¯h V +� +f(w) +� +. +(11.4) +The amplitude reads then +A0,3 = +� 3 +� +i=1 +fi ◦ Vi(0) +� +S2 += +� 3 +� +i=1 +f ′ +i(0)hif ′ +i(0) +¯hi +� � 3 +� +i=1 +Vi +� +fi(0) +� +� +S2 +(11.5a) +∝ +� 3 +� +i=1 +f ′ +i(0)hif ′ +i(0) +¯hi +� +� +f1(0) − f2(0) +�h3−h1−h2 × perms × c.c. +(11.5b) +The amplitude depends on the local coordinate choice fi, but not on the metric around the +punctures. It is also invariant under SL(2, C): the transformation (11.2) written in terms of +the local coordinates is +fi −→ afi + b +cfi + d +(11.6) +from which we get: +f ′ +i −→ +f ′ +i +(cfi + d)2 , +fi − fj −→ +fi − fj +(cfi + d)(cfj + d). +(11.7) +All together, this implies the invariance of the 3-point amplitude since the factors in the +denominator cancel. +When the states are on-shell hi = 0, the dependence in the local +coordinate cancels, showing that the latter is non-physical. +One can ask how Feynman graphs can be constructed in string theory. By definition, an +amplitude is the sum of Feynman graphs contributing at that order in the loop expansion +and for the given number of external legs. The Feynman graphs are themselves built from a +set of Feynman rules. These correspond to the data of the fundamental interactions together +with the definition of a propagator. Since a tree-level cubic interaction is the interaction of +the lowest order, it makes sense to promote it to a fundamental cubic vertex2 V0,3: +V0,3(V1, V2, V3) := += A0,3(V1, V2, V3). +(11.8) +The index 0 reminds that it is a tree-level interaction. +2The notation will become clear later, and should not be confused with the vertex operators. +172 + +11.1.2 +4-point function +The tree-level 4-point amplitude is expressed as +A0,4 = +� +d2z4 +� 3 +� +i=1 +c¯cVi(zi) V4(z4) +� +S2 +. +(11.9) +The conformal weights are denoted by h(Vi) = hi. For on-shell states, hi = 1: while there +is no dependence on the positions z1, z2 and z3, there are divergences for +z4 −→ z1, z2, z3, +(11.10) +corresponding to collisions of punctures in the integration process. Moreover, the expression +does not look symmetric: it would me more satisfactory if all the insertions were accompanied +by ghost insertions and if all the puncture locations were treated on an equal footing. +Example 11.1 – Tachyons +Given tachyon states Vi = eiki·X, the amplitude reads: +A0,4 ∝ +3 +� +i,j=1 +i |q|1/2, +(12.32) +and conversely. The idea is that the disk D(1) +q +of Σg1,n1 is removed and replaced by the +complement of D(2) +q +in Σg2,n2, i.e. the full surface Σg2,n2 −D(2) +q +is glued inside D(1) +q . While it +is clear geometrically, this statement may look confusing from the coordinate point of view +because the local coordinates w(1) +n1 and w(2) +n2 do not cover completely the Riemann surfaces, +but their relation still encodes information about the complete surface. The reason is that +one can always use transition functions to relate the coordinates on the two surfaces. +Example 12.1 +Denote by S(1) +a +and S(2) +b +the spheres sharing a boundary with D(1) +n1 and D(2) +n2 , and write +the corresponding coordinates by z(1) +a +and z(2) +b +such that the transition functions are +z(1) +a += f (1) +an1(w(1) +n1 ), +z(2) +b += f (2) +bn2(w(2) +n2 ). +(12.33) +188 + +(a) Direct gluing of circles. +(b) Connection by a long tube. +(c) Insertion of the second surface into the first one. +(d) Insertion of the first surface +into the second one. +Figure 12.5: Different representations of the surface Σ1,2 obtained after gluing Σ1,1 and Σ0,3 +through the plumbing fixture. +Figure 12.6: Smoothed connection between both surfaces. +189 + +Then the coordinates za and zb are related by +z(1) +a += f (1) +an1 +� +w(1) +n1 +� += f (1) +an1 +� +q +w(2) +n2 +� += f (1) +an1 +� +q +f (2)−1 +bn2 +� +z(2) +b +� +� +(12.34) +such that the new transition function reads +za = Fab(zb), +Fab = f (1) +an1 ◦ (q · I) ◦ f (2)−1 +bn2 +, +(12.35) +where I is the inversion (the superscript on the coordinates za and zb has been removed +to indicate that they are now seen as coordinates on the same surface Σg,n). +The Riemann surface Σg,n is a point of Mg,n. By varying the moduli parameters of +Σg1,n1 and Σg2,n2, one obtains other surfaces in Mg,n. +But the number of parameters +furnished by Σg1,n1 and Σg2,n2 does not match the dimension (11.53) of Mg,n: +Mg1,n1 + Mg2,n2 = 6g1 − 6 + 2n1 + 6g2 − 6 + 2n2 = Mg,n − 2. +(12.36) +This means that the subspace of Mg,n obtained by gluing all the possible surfaces in Mg1,n1 +and Mg2,n2 is of codimension 2. +The missing complex parameter is q: in writing the +plumbing fixture, it was taken to be fixed, but it can be varied to generate a 2-parameter +family of Riemann surfaces in Mg,n, with the moduli of the original surfaces held fixed. +The surface Σg,n is equipped with local coordinates inherited from the original surfaces +Σg1,n1 and Σg2,n2. Hence, the plumbing fixture of points in Pg1,n1 and Pg2,n2 automatically +leads to a point of Pg,n. The fact that the local coordinates are inherited from lower-order +surfaces is called gluing compatibility. It is also not necessary to add parameters to describe +the fibre direction. +12.3.2 +Non-separating case +In the previous section, the plumbing fixture was used to glue punctures on two different +surfaces. In fact, one can also glue two punctures on the same surface to get a new surface +with an additional handle: +Σg,n = #Σg1,n1, +� +g = g1 + 1, +n = n1 − 2, +(12.37) +defining # as a unary operator. This gluing is called non-separating because there is a single +surface before the identification of the disks. +In terms of the local coordinates, the gluing relation reads +w(1) +n1−1w(1) +n1 = q, +(12.38) +where we consider the last two punctures for definiteness. +The dimensions of both moduli spaces are related by +Mg1,n1 = Mg,n − 2. +(12.39) +Again, the two missing parameters are provided by varying q and we obtain a Mg,n- +dimensional subspace of Mg,n. +Example 12.2 +Here are some examples of surfaces obtained by gluing: +190 + +• Σ0,4 = Σ0,3#Σ0,3 +• Σ0,5 = Σ0,3#Σ0,3#Σ0,3, Σ0,3#Σ0,4 +• Σ1,1 = #Σ0,3 +• Σ1,2 = #Σ0,4, Σ1,1#Σ0,3 +Note that the moduli on the LHS and RHS are fixed (we will see later that not all +surfaces can be obtained by gluing). +12.3.3 +Decomposition of moduli spaces and degeneration limit +We have seen that the separating and non-separating plumbing fixtures yield a family of +surfaces in Mg,n described in terms of lower-dimensional moduli spaces. The question is +whether all points in Mg,n can be obtained in this way by looking at all the possible gluing +(varying g1, n1, g2 and n2). It turns out that this is not possible, which is at the core of the +difficulties to construct a string field theory. +Which surfaces are obtained from this construction? In order to interpret the regions +of Mg,n covered by the plumbing fixture, the parametrization (12.31) is the most useful. +Previously, we explained that s gives the size of the tube connecting the two surfaces. Since +the latter is like a sphere with two punctures, it corresponds to a cylinder (interpreted as an +intermediate closed string propagating). The angle θ in (12.31) is the twist of the cylinder +connecting both components. This amounts to start with θ = 0, then to cut the cylinder, +to twist it by an angle θ and to glue again. +The limit s → ∞ (|q| → 0) is called the degeneration limit: the degenerate surface +Σg,n reduces to Σg1,n1 and Σg2,n2 connected by a very long tube attached to two punctures +(separating case), or to Σg−1,n+2 with a very long handle (non-separating case). So it means +that the family of surfaces described by the plumbing fixture are “close” to degeneration. +Another characterization (for the separating case) is that the punctures on Σg1,n1 are closer +(according to some distance, possibly after a conformal transformation) to each other than +to the punctures on Σg2,n2. +Conversely, there are surfaces which cannot be described in this way: the plumbing +fixture does not cover all the possible values of the moduli. For a given Mg,n, we denote +the surfaces which cannot be obtained by the plumbing fixture by Vg,n. This space does +not contain any surface arbitrarily close to degeneration (i.e. with long handles or tubes). +In terms of punctures, it also means that there is no conformal frame where the punctures +split in two sets. +In the previous subsection, we considered two specific punctures, but any other punctures +could be chosen. Hence, there are many ways to split Σg,n in two surfaces Σg1,n1 and Σg2,n2 +(with fixed g1, g2, n1 and n2): every partition of the punctures and holes in two sets lead to +different degeneration limits (because they are associated to different moduli – Figure 12.7). +Since each puncture is described by a modulus, choosing different punctures for gluing give +different set of moduli for Σg,n, such that each possibility covers a different subspace of +Mg,n. The part of the moduli space Mg,n covered by the plumbing fixture of all surfaces +Σg1,n1 and Σg2,n2 (with fixed g1, g2, n1, n2) is denoted by Mg1,n1#Mg2,n2: +Mg1,n1#Mg2,n2 ⊂ Mg,n, +(12.40) +where the operation # includes the plumbing fixture for all values of q and all pairs of +punctures. Similarly, the part covered by the non-separating plumbing fixture is written as +#Mg1,n1: +#Mg1,n1 ⊂ Mg,n. +(12.41) +Importantly, the regions covered by the plumbing fixture depend on the choice of the +local coordinates because (12.30) is written in terms of local coordinates. The subspaces +Mg1,n1#Mg2,n2 and #Mg1,n1 are not necessarily connected (in the topological sense). +191 + +(a) Degeneration 12 → 34 +(b) Degeneration 13 → 24 +(c) Degeneration 14 → 23 +Figure 12.7: Permutations of punctures while gluing two spheres: they correspond to differ- +ent (disconnected) parts of M0,4. +The moduli space Mg,n cannot be completely covered by the plumbing fixture of lower- +dimensional surfaces. We define the propagator and fundamental vertex regions Fg,n and +Vg,n as the subspaces which can and cannot be described by the plumbing fixture: +Fg,n := #Mg−1,n+2 +� +� +� +n1+n2=n+2 +g1+g2=g +Mg1,n1#Mg2,n2 +� +, +(12.42a) +Vg,n := Mg,n − Fg,n, +(12.42b) +In the RHS, it is not necessary to consider multiple non-separating plumbing fixtures for +the first term because #Mg−2,n+4 ⊂ Mg−1,n+2, etc. For the same reason, it is sufficient to +consider a single separating plumbing fixture. Note that Vg,n and Fg,n are in general not +connected subspaces. A simple illustration is given in Figure 12.9. The actual decomposition +of M0,4 is given in Figure 12.8. Importantly, Fg,n and Vg,n depend on the choice of the +local coordinates for all Vg′,n′ appearing in the RHS. +It is also useful to define the subspaces F1PR +g,n +and V1PI +g,n of Mg,n which can and cannot +be described with the separating plumbing fixture only: +F1PR +g,n := +� +n1+n2=n+2 +g1+g2=g +Mg1,n1#Mg2,n2, +(12.43a) +V1PI +g,n := Mg,n − F1PR +g,n . +(12.43b) +1PR (1PI) stands for 1-particle (ir)reducible, a terminology which will become clear later. +Note the relation: +V1PI +g,n = Vg,n +� +� � +g′ +#Mg−g′,n+g′ +� +. +(12.44) +The two plumbing fixtures behave as follow: +• separating: increases both n and g (if both surfaces have a non-vanishing g); +192 + +Figure 12.8: In white are the subspaces of the moduli space M0,4 covered by the plumb- +ing fixture. The three different regions correspond to the three different ways to pair the +punctures (see Figure 12.7). In grey is the fundamental vertex region V0,4. +Figure 12.9: Schematic illustration of the covering of Mg,n from the plumbing fixture of +lower-dimensional spaces. The fundamental region Vg,n (usually disconnected) is not covered +by the plumbing fixture. +193 + +• non-separating plumbing: increases g but decreases n. +The construction is obviously recursive: starting from the lowest-dimensional moduli space, +which is M0,3 (no moduli), one has: +V0,3 = M0,3, +F0,3 = ∅. +(12.45) +Next, the subspace of M0,4 obtained from the plumbing fixture is: +F0,4 = V0,3#V0,3, +(12.46) +and V0,4 is characterized as the remaining region. Then, one has: +F0,5 = M0,4#M0,3 += F0,4#V0,3 + V0,4#V0,3 = V0,3#V0,3#V0,3 + V0,4#V0,3, +(12.47) +and V0,5 is what remains of M0,5. The pattern continues for g = 0. The same story holds +for g ≥ 1: the first such space is +F1,1 = #V0,3, +(12.48) +and V1,1 = M1,1 − F1,1. The gluing of a 3-punctured sphere and the addition of a handle +are the two most elementary operations. +To keep track of which moduli spaces can contribute, it is useful to find a function of +Σg,n, called the index, which increases by 1 for each of the two elementary operations: +r(Σg1,n1#Σ0,3) = r(Σg1,n1) + 1, +r(#Σg1,n1) = r(Σg1,n1) + 1. +(12.49) +An appropriate function is +r(Σg,n) = 3g + n − 2 ∈ N∗. +(12.50) +which is normalized such that: +r(Σ0,3) = 1. +(12.51) +For a generic separating plumbing fixture, we find: +r(Σg1,n1#Σg2,n2) = r(Σg1,n1) + r(Σg2,n2). +(12.52) +Since the index increases, surfaces with a given r can be obtained by considering all the +gluings of surfaces with r′ < r. +12.3.4 +Stubs +To conclude this chapter, we introduce the concept of stubs. Previously in (12.31), the range +of the parameter s was the complete line of positive numbers, s ∈ R+. This means that +tubes of all lengths were considered to glue surfaces. But, we could also introduce a minimal +length s0 > 0, called the stub parameter, for the tube. In this case, the plumbing fixture +parameter is generalized to: +q = e−s+iθ, +s ∈ [s0, ∞), +θ ∈ [0, 2π), +s0 ≥ 0. +(12.53) +What is the effect on the subspaces Fg,n(s0) and Vg,n(s0)? Obviously, less surfaces can be +described by the plumbing fixture if s0 > 0 than if s0 = 0, since the plumbing fixture cannot +describe anymore surfaces which contain a tube of length less than s0. Equivalently, the +values of the moduli described by the plumbing fixture is more restricted when s0 > 0. More +generally, one has: +s0 < s′ +0 : +Fg,n(s′ +0) ⊂ Fg,n(s0) +Vg,n(s0) ⊂ Vg,n(s′ +0). +(12.54) +194 + +Figure 12.10: In light grey is the subspace covered by the V0,4(s0) as in Figure 12.8. In dark +grey is the difference δV0,4 = V0,4(s0 + δs0) − V0,4(s0) with δs0 > 0. +This is illustrated on Figure 12.10. Even if s0 is very large, Vg,n still does not include surfaces +arbitrarily close to degeneracy. In general, we omit the dependence in s0 except when it is +necessary. +To interpret the stub parameter, consider two local coordinates w1 and w2 and rescale +them by λ ∈ C with Re λ > 0: +w1 = λ ˜w1, +w2 = λ ˜w2. +(12.55) +Then, the plumbing fixture (12.30) becomes +˜w1 ˜w2 = e−˜s+i˜θ. +(12.56) +with +˜s = s + 2 ln |λ|, +˜θ = θ + i ln λ +¯λ. +(12.57) +If s ∈ R+, the corresponding range of ˜s is +˜s ∈ [s0, ∞), +s0 := 2 ln |λ|. +(12.58) +This shows that rescaling the local coordinates by a constant parameter is equivalent to +change the stub parameter. +Note also how performing a global phase rotation in (12.57) is equivalent to shift the +twist parameter. Working in ˆPg,n forces to take λ ∈ R+. +12.4 +Summary +In this chapter, we have explained how to parametrize the fibre bundle Pg,n, that is, appro- +priate coordinates for the moduli space and the local coordinate systems. This was realized +195 + +by introducing different coordinate patches and encoding all the informations of Pg,n in +the transition functions. Then, this description lead to a simple description of the tangent +vectors through the Schiffer variation. +In the next chapter, we will continue the program by building the p-forms required to +describe off-shell amplitudes. +12.5 +Suggested readings +• Plumbing fixture [193, sec. 9.3]. +196 + +Chapter 13 +Off-shell amplitudes +Abstract +While the previous chapter was purely geometrical, this one makes contact +with string theory through the worldsheet CFT. We continue the description of Pg,n by +constructing p-forms. +The reason why we need to consider the CFT is that ghosts are +necessary to build the p-forms: this can be understood from Chapter 2, where we found +that the ghosts must be interpreted as part of the measure on the moduli space. Then, we +build the off-shell amplitudes and discuss some properties. +13.1 +Cotangent spaces and amplitudes +In this section, we construct the p-forms on Pg,n which are needed for the amplitudes. We +first motivate the expressions from general ideas, and check later that they have the correct +properties. +13.1.1 +Construction of forms +A p-form ω(g,n) +p +∈ �p T ∗Pg,n is a multilinear antisymmetric map from �p TPg,n to a function +of the moduli parameters. The superscript on the form is omitted when there is no ambiguity +about the space considered. The components ωi1···ip of the p-form are defined by inserting +p basis vectors ∂s1, . . . , ∂sp +ωi1···ip := ωp(∂s1, . . . , ∂sp), +(13.1) +where ∂s = +∂ +∂xs and xs are the coordinates (12.13). It is antisymmetric in any pair of two +indices +ωi1i2···ip = −ωi2i1···ip, +(13.2) +and multilinearity implies that +ωp +� +V (1), . . . , V (p)� += ωp +� +V (1) +s1 ∂s1, . . . , V (p) +sp ∂sp +� += ωi1···ipV (1) +s1 · · · V (p) +sp , +(13.3) +given vectors V (α) = V (α) +s +∂s. +The p-forms which are needed to define off-shell amplitudes depend on the external states +Vi (i = 1, . . . , n) inserted at the punctures zi. They are maps from �p TPg,n × Hn to a +function on Pg,n. The dependence on the states is denoted equivalently as +ωp(V1, . . . , Vn) := ωp(⊗iVi). +(13.4) +The simplest way to get a function on Pg,n from the states Vi is to compute a CFT correlation +function of the operators inserted at the points zi = fi(0) on the surface Σg,n described by +the point in Mg,n. +197 + +The 0-form is just a function and is defined by: +ω0 = (2πi)−Mc +g,n +� n +� +i=1 +fi ◦ Vi(0) +� +Σg,n +. +(13.5) +For simplicity, the dependence in the local coordinates fi is kept implicit in the rest of the +chapter. +A natural approach for constructing p-forms is to build them from elementary 1-forms +and to use ghosts to enforce the antisymmetry. +Remembering the Beltrami differentials +found in Chapter 2, the contour integral of ghosts b(z) weighted by some vector field is a +good starting point. In the current language, it is defined by its contraction with a vector +V = (v, C) ∈ TPg,n defined in (12.23): +B(V ) := +� +C +dz +2πi b(z)v(z) + +� +C +d¯z +2πi +¯b(¯z)¯v(¯z), +(13.6) +where b(z) and ¯b(¯z) are the b-ghost components, and v is the vector field on Σg,n defining +V . The contours run anti-clockwise. If the contour C includes several circles (C = ∪αCα), +B(V ) is defined as the sum of the contour integral on each circle: +B(V ) := +� +α +� +Cα +dz +2πi b(z)v(z) + c.c. +(13.7) +It is also useful to define another object built from the energy–momentum tensor: +T(V ) := +� +C +dz +2πi T(z)v(z) + +� +C +d¯z +2πi +¯T(¯z)¯v(¯z), +(13.8) +where T and ¯T are the components of the energy–momentum tensor. It is defined such that +T(V ) = {QB, B(V )}. +(13.9) +Considering the coordinate system (12.13), the Beltrami form can be decomposed as: +B = Bsdxs, +Bs := B(∂s), +(13.10a) +Bs = +� +α +� +Cα +dσα +2πi b(σα) ∂Fα +∂xs +� +F −1 +α (σα) +� ++ +� +α +� +Cα +d¯σα +2πi +¯b(¯σα) ∂ ¯Fα +∂xs +� ¯F −1 +α (¯σα) +� +, +(13.10b) +where the contour orientations are defined by having the σα coordinate system on the left. +We define the p-form contracted with a set of vectors V (1), . . . , V (p) by +ωp +� +V (1), . . . , V (p)� +(V1, . . . , Vn) := (2πi)−Mc +g,n +� +B(V (1)) · · · B(V (p)) +n +� +i=1 +Vi +� +Σg,n +, +(13.11) +and the corresponding p-form reads +ωp = ωp,s1···sp dxs1 ∧ · · · ∧ dxsp +(13.12a) += (2πi)−Mc +g,n +� +Bs1dxs1 ∧ · · · ∧ Bspdxsp +n +� +i=1 +Vi +� +Σg,n +. +(13.12b) +In this expression, the form contains an infinite numbers of components ωp,s1···sp since there +is an infinite number of coordinates. Note that the normalization is independent of p. +198 + +In practice, one is not interested in Pg,n, but rather in a subspace of it. Given a q- +dimensional subspace S of Pg,n parametrized by q real coordinates t1, . . . , tq +xs = xs(t1, . . . , tq), +(13.13) +the restriction of a p-form to this subspace is obtained by the chain rule: +∀p ≤ q : +ωp|S = (2πi)−Mc +g,n +� +Br1 +∂xs1 +∂tr1 +dtr1 ∧ · · · ∧ Brp +∂xsp +∂trp +dtrp +n +� +i=1 +Vi +� +Σg,n +, +∀p > q : +ωp|S = 0. +(13.14) +We will often write the expression directly in terms of the coordinates of S and abbreviate +the notation as: +Br := ∂xs +∂tr +Bs. +(13.15) +13.1.2 +Amplitudes and surface states +It is now possible to write the amplitude more explicitly. An on-shell amplitude is defined +as an integral over Mg,n. Off-shell, one needs to consider local coordinates around each +puncture, that is, a point of the fibre for each point of the base Mg,n. This defines a Mg,n- +dimensional section Sg,n of Pg,n (Figure 11.2). The g-loop n-point off-shell amplitude of the +states V1, . . . , Vn reads: +Ag,n(V1, . . . , Vn)Sg,n := +� +Sg,n +ωg,n +Mg,n(V1, . . . , Vn) +�� +Sg,n, +(13.16a) +ωg,n +Mg,n(V1, . . . , Vn) +�� +Sg,n = (2πi)−Mc +g,n +�Mg,n +� +λ=1 +Bs +∂xs +∂tλ +dtλ +n +� +i=1 +fi ◦ Vi(0) +� +Σg,n +, +(13.16b) +where the choice of the fi is dictated by the section Sg,n. From now on, we stop to write the +restriction of the form to the section. We also restrict to the cases where χg,n = 2−2g−n < 0. +The complete (perturbative) n-point amplitude is the sum of contributions from all loops: +An(V1, . . . , Vn) := +� +g≥0 +Ag,n(V1, . . . , Vn). +(13.17) +More generally, we define the integral over a section Rg,n which projection on the base +is a subspace of Mg,n (and not the full space as for the amplitude) as: +Rg,n(V1, . . . , Vn) := +� +Rg,n +ωg,n +Mg,n(V1, . . . , Vn), +(13.18) +For simplicity, we will sometimes use the same notation for the section of Pg,n and its +projection on the base Mg,n. For this reason, the reader should assume that some choice of +local coordinates around the punctures is made except otherwise stated. +Given sections Rg,n, the sum over all genus contribution is written formally as +Rn := +� +g≥0 +Rg,n, +(13.19) +such that +Rn(V1, . . . , Vn) := +� +g≥0 +Rg,n(V1, . . . , Vn) = +� +g≥0 +� +Rg,n +ωg,n +Mg,n(V1, . . . , Vn). +(13.20) +199 + +A surface state is defined as a n-fold bra which reproduces the expression of a given +function when contracted with n states Ai. The surface ⟨Σg,n|, form ⟨ωg,n|, section ⟨Rg,n| +and amplitude ⟨Ag,n| n-fold states are defined by the following expressions: +⟨Σg,n| Bs1 · · · Bsp | ⊗i Vi⟩ := ωs1···sp(V1, . . . , Vn), +(13.21a) +⟨ωg,n +p +| ⊗i Vi⟩ := ωp(V1, . . . , Vn), +(13.21b) +⟨Ag,n| ⊗i Vi⟩ := Ag,n(V1, . . . , Vn). +(13.21c) +The last relation is generalized to any section Rg,n: +⟨Rg,n| ⊗i Vi⟩ := Rg,n(V1, . . . , Vn). +(13.21d) +The reason for introducing these objects is that the form (13.12) is a linear map from H⊗n +to a form on Mg,n – see (13.4). Thus, there is always a state ⟨Σg,n| such that its BPZ +product with the states reproduces the form. In particular, the state ⟨Σg,n| contains all the +information about the local coordinates and the moduli (the dependence is kept implicit). +The definition of the other states follow similarly. These states are defined as bras, but they +can be mapped to kets. +One finds the obvious relations: +⟨ωg,n +p +| =⟨Σg,n| Bs1dxs1 · · · Bspdxsp, +⟨Ag,n| = +� +Mg,n +⟨ωg,n +p +| . +(13.22) +The surface states don’t contain information about the matter CFT: they collect the +universal data (like local coordinates) needed to describe amplitudes. Hence, it is an im- +portant step in the description of off-shell string theory to characterize this data. However, +note that the relation between a surface state and the corresponding form does depend on +the CFT. +Example 13.1 – On-shell amplitude A0,4 +The transition functions are given by (see Figure 12.1): +C1 : w1 = z1 − y1, +C3 : w3 = z2 − y3, +C5 : z1 = z2, +C2 : w2 = z1 − y2, +C4 : w4 = z2 − y4. +(13.23) +Three of the parameters (y1, y2 and y3) are fixed while the single complex modulus +of M0,4 is taken to be y4. Since we are interested in the on-shell amplitude, it is not +necessary to introduce local coordinates and the associated parameters. +A variation of the modulus +y4 −→ y4 + δy4, +¯y4 −→ ¯y4 + δ¯y4 +(13.24) +is equivalent to a change in the transition function of C4. This translates in turn into +a transformation of z2: +z′ +2 = z2 + δy4, +¯z′ +2 = ¯z2 + δ¯y4. +(13.25) +Then, the tangent vector V = ∂y4 is associated to the vector field +v = 1, +¯v = 0, +(13.26) +with support on C4. For V = ∂¯y4, one finds +v = 0, +¯v = 1. +(13.27) +200 + +The Beltrami 1-form for the unit vectors are +B(∂y4) = +� +C4 +dz2 b(z2)(+1), +B(∂¯y4) = +� +C4 +d¯z2 ¯b(¯z2)(+1), +(13.28) +with both contours running anti-clockwise. +The components of the 2-form reads +ω2(∂y4, ∂¯y4) = +1 +2πi +� +B(∂y4)B(∂¯y4) +4 +� +i=1 +Vi +� +Σ0,4 += +1 +2πi +�� +C4 +dz2 b(z2) +� +C4 +d¯z2 ¯b(¯z2) +4 +� +i=1 +Vi +� +Σ0,4 +. +For on-shell states Vi = c¯cVi(yi, ¯yi), this becomes +ω2(∂y4, ∂¯y4) = +1 +2πi +� 3 +� +i=1 +c¯cVi(yi, ¯yi) +� +C4 +dz2 b(z2) +� +C4 +d¯z2 ¯b(¯z2)¯c(¯y4)c(y4)V4(y4, ¯y4) +� +Σ0,4 +. +The first three operators could be moved to the left because they are not encircled by the +integration contour. Note the difference with the example discussed in Section 11.1.2: +here, the contour encircles z3, while it was encircling y3 for the s-channel. +Using the OPE +� +C4 +dz2 b(z2)c(y4) ∼ +� +C4 +dz2 +1 +z2 − y4 +(13.29) +to simplify the product of b and c gives the amplitude +A0,4 = +1 +2πi +� +dy4 ∧ d¯y4 +� 3 +� +i=1 +c¯cVi(yi) V4(y4) +� +Σ0,4 +. +(13.30) +This is the standard formula for the 4-point function derived from the Polyakov path +integral. +13.2 +Properties of forms +In this section, we check that the form (13.12) has the correct properties: +• antisymmetry under exchange of two vectors; +• given a trivial vector of (a subspace of) Pg,n (Section 12.1), its contraction with the +form vanishes: ωp(V (1), . . . , V (p)) = 0 if any of the V (i) generates: +– reparametrizations of za for V (i) ∈ TPg,n, +– rotation wi → (1 + iαi)wi for V (i) ∈ T ˆPg,n, +– reparametrizations of wi keeping wi = 0 if the states are on-shell for V (i) ∈ +TMg,n; +• BRST identity, which is necessary to prove several properties of the amplitudes. +The first property is obvious. Indeed, the form is correctly antisymmetric under the +exchange of two vectors V (i) and V (j) due to the ghost insertions. +201 + +13.2.1 +Vanishing of forms with trivial vectors +Reparametrization of za +Consider the sphere Sa with coordinate za, and denote by C1, +C2 and C3 the three boundaries. Then, a reparametrization +za −→ za + φ(za) +(13.31) +is generated by a vector field φ(z) which is regular on Sa. This transformation modifies +the transition functions on the three circles and is thus associated to a tangent vector V +described by a vector field v with support on the three circles: +Ci : +v(i) = φ|Ci. +(13.32) +The Beltrami form then reads +B(V ) = +3 +� +i=1 +� +Ci +dza b(za)φ(za) + c.c. +(13.33) +where the orientations of the contours are such that Sa is on the left. Since the vector +field φ is regular in Sa, two of the contours can be deformed until they merge together. +The resulting orientation is opposite to the one of the last contour (Figure 13.1). As a +consequence, both cancel and the integral vanishes. +Figure 13.1: Deformation of the contour of integration defining the Beltrami form for a +reparametrization of za. +The figure is drawn for two circles at a hole, but the proof is +identical for other types of circles. +Rotation of wi +Consider an infinitesimal phase rotation of the local coordinate wi in the +disk Di: +wi −→ (1 + iαi)wi, +¯wi −→ (1 − iαi) ¯wi, +(13.34) +with αi ∈ R. The tangent vector is defined by the circle Ci and the vector field by +v = iwi, +¯v = −i ¯wi. +(13.35) +The Beltrami form for this vector is +B(V ) = i +� +Ci +dwi wi b(wi) − i +� +Ci +d ¯wi ¯wi ¯b( ¯wi), +(13.36) +where Di is kept to the left. +In the p-form (13.11), the ith operator Vi is inserted in Di and encircled by Ci. Because +there is no other operator inside Di, the contribution of this disk to the form is +B(V )Vi(0) = i +� +Ci +dwi wi b(wi)Vi(0) − i +� +Ci +d ¯wi ¯wi ¯b( ¯wi)Vi(0), +(13.37) +202 + +The state–operator correspondence allows to rewrite this result as +i(b0 − ¯b0) |Vi⟩ , +(13.38) +since the contour integral picks the zero-modes of b and of ¯b. +Requiring that the form +vanishes implies the ghost counter-part of the level-matching condition: +b− +0 |Vi⟩ = 0. +(13.39) +Hence, consistency of off-shell amplitudes imply that +Vi ∈ H−, +(13.40) +where H− is defined in (11.38). +Reparametrization of wi +A reparametrization of the local coordinate wi keeping the +origin of Di fixed reads: +wi −→ f(wi), +f(0) = 0. +(13.41) +The function can be expanded in series: +f(wi) = +� +m≥0 +pmwm+1 +i +. +(13.42) +Because the transformation is holomorphic, it can be extended on Ci. Each parameter pm +provides a coordinate of Pg,n and whose deformation corresponds to a vector field: +vm = wm+1 +i +, +¯vm = 0. +(13.43) +The corresponding Beltrami differential is +B(∂pm) = +� +Ci +dwi b(wi)wm+1 +i +. +(13.44) +Since only the operator Vi is inserted in the disk, the state–operator correspondence gives +bm |Vi⟩. Requiring that the form vanishes on Mg,n for all m and also for the anti-holomorphic +vectors gives the conditions: +∀m ≥ 0 : +bm |Vi⟩ = 0, +¯bm |Vi⟩ = 0. +(13.45) +This holds automatically for on-shell states Vi = c¯cVi. +13.2.2 +BRST identity +The BRST identity for the p-form (13.12) reads +ωp +� � +i +Q(i) +B ⊗i Vi +� += (−1)pdωp−1(⊗Vi), +(13.46) +using the notation (13.4). The BRST operator acting on the ith Hilbert space is written as +Q(i) +B = 1i−1 ⊗ QB ⊗ 1n−i +(13.47) +and acts as +QBVi(z, ¯z) = +1 +2πi +� +dw jB(w)Vi(z, ¯z) + c.c. +(13.48) +203 + +More explicitly, the LHS corresponds to +ωp +� � +i +Q(i) +B ⊗i Vi +� += ωp(QBV1, V2, . . . , Vn) + (−1)|V1|ωp(V1, QBV2, . . . , Vn) ++ · · · + (−1)|V1|+···+|Vn−1|ωp(V1, V2, . . . , QBVn). +(13.49) +We give just an hint of this identity, the complete proof can be found in [262, pp. 85–89, +215, sec. 2.5]. +The contour of the BRST current around each puncture can be deformed, picking singu- +larities due to the presence of the Beltrami forms. Using (13.9), we find that anti-commuting +the BRST charge with the Beltrami form Bs leads to an insertion of +Ts = {QB, Bs}. +(13.50) +The energy–momentum tensor generates changes of coordinates. Hence, Ts = T∂s is precisely +the generator associated to an infinitesimal change of the coordinate xs on Pg,n. The latter +is given by the vector ∂s. For this reason, one can write: +dxs {QB, Bs} = dxs Ts = dxs ∂s = d, +(13.51) +where d is the exterior derivative on Pg,n. The minus signs arise if the states Vi are Grass- +mann odd. +13.3 +Properties of amplitudes +In order for the p-form (13.12) to be non-vanishing, its total ghost number should match +the ghost number anomaly: +Ngh +� +ωp(V1, . . . , Vn) +� += +n +� +i=1 +Ngh(Vi) − p = 6 − 6g, +(13.52) +using Ngh(B) = −1. For an amplitude, one has p = Mg,n = 6g − 6 + 2n and thus: +Ngh(ωMg,n) = 6 − 6g +=⇒ +n +� +i=1 +Ngh(Vi) = 2n. +(13.53) +This condition holds automatically for on-shell states since Ngh(c¯cVi) = 2. +13.3.1 +Restriction to ˆPg,n +The goal of this section is to explain why amplitudes must be described in terms of a section +of ˆPg,n (12.10) instead of Pg,n. +This means that one should identify local coordinates +differing by a global phase rotation. +The off-shell amplitudes (13.16) are multi-valued on Pg,n. Indeed, the amplitude depends +on the local coordinates1 and changes by a factor under a global phase rotation of any local +coordinate wi → eiαwi. However, such a global rotation leaves the surface unchanged, since +the flat metric |dwi|2 is invariant. This means that the same surface leads to different values +for the amplitude. To prevent this multi-valuedness of the amplitudes, it is necessary to +identify local coordinates differing by a constant phase. +A second way to obtain this condition is to require that the section Sg,n is globally +defined: every point of the section should correspond to a single point of the moduli space +1The current argument does not apply for on-shell amplitudes. +204 + +Mg,n. However, there is a topological obstruction which prevents finding a global section in +Pg,n in general. One hint [56, sec. 2, 71, sec. 3] is to exhibit a nowhere vanishing 1-form if +Sg,n is globally defined: this leads to a contradiction since such a 1-form does not generally +exist (see for example [59, sec. 6.3.2, ch. 7]). Then, consider a closed curve in the moduli +space (such curves exist since Mg,n is compact). Starting at a given point Σ of the curve, +one finds that the local coordinates typically change by a global phase when coming back +to the point Σ (Figure 13.2), since this describes the same surface and there is no reason +to expect the phase to be invariant. Up to this identification, it is possible to find a global +section. The latter corresponds to a section of ˆPg,n. +(a) Closed curve in Mg,n. +(b) Change in the phase of wi. +Figure 13.2: Schematic plot of the change in the phase of the local coordinate wi as one +follows a closed curve in Mg,n. If the original phase at Σ is α0 and if the phase varies +continuously along the path, then α1 ̸= α0 when returning back to Σ by continuity. +Remark 13.1 (Degeneracy of the antibracket) It is possible to define a BV structure +on Riemann surfaces [233, 234]. The antibracket is degenerate in Pg,n but not in ˆPg,n [234]. +Global phase rotations of the local coordinates are generated by L− +0 . Hence, identifying +the local coordinates wi → eiαiwi amounts to require that the amplitude is invariant under +L− +0 . This is equivalent to imposing the level-matching condition +L− +0 |Vi⟩ = 0 +(13.54) +on the off-shell states. This condition was interpreted in Section 3.2.2 as a gauge-fixing +condition for translations along the S1 of the string. This shows, in agreement with earlier +comments, that the level-matching condition should also be imposed off-shell because no +gauge symmetry is introduced for the corresponding transformation. +If the generator L− +0 is trivial, this means that the ghost associated to the corresponding +tangent vector must be decoupled. According to Section 13.2.1, this corresponds to the +constraint: +b− +0 |Vi⟩ = 0. +(13.55) +This can be interpreted as a gauge fixing condition (Section 10.5), which could in principle +be relaxed. However, the decoupling of physical states (equivalent to gauge invariance in +SFT) happens only after integrating over the moduli space. This requires having a globally +defined section. +As a consequence, off-shell states are elements of the semi-relative Hilbert space +Vi ∈ H− ∩ ker L− +0 , +(13.56) +and the amplitudes are defined by integrating the form ωMg,n over a section Sg,n ⊂ ˆPg,n. +205 + +Computation – Equation (13.54) +The operator associated to the state through |Ai⟩ = Ai(0) |0⟩ transforms as +Vi(0) −→ (eiαi)h(e−iαi) +¯hVi(0) +(13.57) +which translates into +|Vi⟩ −→ eiαi(L0−¯L0) |Vi⟩ +(13.58) +for the state, using the fact that the vacuum is invariant under L0 and ¯L0. +Then, +requiring the invariance of the state leads to (13.54). +13.3.2 +Consequences of the BRST identity +Two important properties of the on-shell amplitudes can be deduced from the BRST identity +(13.46): the independence of physical results on the choice of local coordinates and the +decoupling of pure gauge states. +Given BRST closed states, the LHS of (13.46) vanishes identically +∀i : +QB |Vi⟩ = 0 +=⇒ +dωp−1(V1, . . . , Vn) = 0. +(13.59) +Using this result, one can compare the on-shell amplitudes computed for two different sec- +tions S and S′: +� +S +ωMg,n − +� +S′ ωMg,n = +� +∂T +ωMg,n−1 = +� +T +dωMg,n−1 = 0, +(13.60) +using Stokes’ theorem and where T is the surface delimited by the two sections (Figure 13.3). +This implies that on-shell amplitudes do not depend on the section, and thus on the local +coordinates. In obtaining the result, one needs to assume that the vertical segments do +not contribute. The latter correspond to boundary contributions of the moduli space. In +general, many statements hold up to this condition, which we will not comment more in this +book. +Figure 13.3: Two sections S and S′ of Pg,n delimiting a surface T . +Next, we consider a pure gauge state together with BRST closed states: +|V1⟩ = QB |Λ⟩ , +QB |Vi⟩ = 0. +(13.61) +The BRST identity (13.46) reads: +ωMg,n(QBΛ, V2, . . . , Vn) = dωMg,n−1(Λ, V2, . . . , Vn), +(13.62) +206 + +which gives the amplitude +� +S +ωMg,n(QBΛ, V2, . . . , Vn) = +� +S +dωMg,n−1(Λ, V2, . . . , Vn) = +� +∂S +ωMg,n−1(Λ, V2, . . . , Vn) +(13.63) +where the last equality follows from Stokes’ theorem. +Assuming again that there is no +boundary contribution, this vanishes: +� +S +ωMg,n(QBΛ, V2, . . . , Vn) = 0. +(13.64) +This implies that pure gauge states decouple from the physical states. +13.4 +Suggested readings +• Definition of the forms [71, 73, 215, 262]. +• Global phase rotation of local coordinates [56, sec. 2, 71, sec. 3, 174, 262, p. 54]. +207 + +Chapter 14 +Amplitude factorization and +Feynman diagrams +Abstract +In the previous chapter, we built the off-shell amplitudes by integrating forms +on sections of Pg,n. Studying their factorizations lead to rewrite them in terms of Feynman +diagrams, which allows to identify the fundamental interactions vertices. We will then be +able to write the SFT action in the next chapter. +14.1 +Amplitude factorization +We have seen how to write off-shell amplitudes. The next step is to rewrite them as a sum +of Feynman diagrams through factorization of amplitudes. +Factorization consists in writing a g-loop n-point amplitude in terms of lower-order +amplitudes in both g and n connected by propagators. Since an amplitude corresponds +to a sum over all possible processes, which corresponds to integrating over the moduli space, +it is natural to associate Feynman diagrams to different subspaces of the moduli space. One +can expect that the plumbing fixture (Section 12.3) is the appropriate translation of the +factorization at the level of Riemann surfaces. We will assume that it is the case and check +that it is correct a posteriori. +To proceed, we consider the contribution to the amplitude Ag,n of the family of surfaces +obtained by the plumbing fixture of two surfaces (separating case) or a surface with itself +(non-separating case). +14.1.1 +Separating case +In this section, we consider the separating plumbing fixture where part of the moduli space +Mg,n is covered by Mg1,n1#Mg2,n2 with g = g1 + g2 and n = n1 + n2 − 2 (Section 12.3.1). +The local coordinates read w(1) +i +and w(2) +j +for i = 1, . . . , n1 and j = 1, . . . , n2. By convention, +the last coordinate of each set is used for the plumbing fixture: +w(1) +n1 w(2) +n2 = q. +(14.1) +The g-loop n-point amplitude with external states {V (1) +1 +, . . . , V (1) +n1−1, V (2) +1 +, . . . , V (2) +n2−1} is +denoted as: +Ag,n = +� +Sg,n +ωg,n +Mg,n +� +V (1) +1 +, . . . , V (1) +n1−1, V (2) +1 +, . . . , V (2) +n2−1 +� +. +(14.2) +208 + +We need to study the form ωg,n +Mg,n on Mg1,n1#Mg2,n2, which means to rewrite it in terms +of the data from Mg1,n1 and from Mg2,n2. This corresponds to the degeneration limit where +the two groups of punctures denoted by V (1) +i +and V (2) +j +(i = 1, . . . , n1 − 1, j = 1, . . . , n2 − 1) +together with g1 and g2 holes move apart from each other (Figure 14.1). +Figure 14.1: Degeneration limit of Σg,n where the punctures V (1) +i +and V (2) +j +move apart from +each other. +Since q is a coordinate of Pg,n, its variation is associated with a tangent vector and a +Beltrami 1-form. The latter has to be inserted inside ωg,n +Mg,n. A change q → q + δq translates +into a change of coordinate +w′(1) +n1 = w(1) +n1 + w(1) +n1 +q +δq, +(14.3) +where w(2) +n2 is kept fixed (obviously, this choice is conventional as explained in Section 12.2). +Thus, the vector field and the Beltrami form are +vq = w(1) +n1 +q +, +Bq = 1 +q +� +Cq +dw(1) +n1 b +� +w(1) +n1 +� +w(1) +n1 . +(14.4) +Computation – Equation (14.3) +Starting from (14.1), vary q → q + δq while keeping w(2) +n2 fixed: +w′(1) +n1 w(2) +n2 = q + δq +w′(1) +n1 = w(1) +n1 +q +(q + δq) = w(1) +n1 + +q +w(1) +n1 +δq. +The second line follows by replacing w(2) +n2 using (14.1). +The Mg,n-form for the moduli described by the plumbing fixture can be expressed as: +ωMg,n +� +V (1) +1 +, . . . , V (1) +Mg1,n1, ∂q, ∂¯q, V (2) +1 +, . . . , V (2) +Mg2,n2 +� +(14.5) += (2πi)−Mc +g,n +�Mg1,n1 +� +λ=1 +B +� +V (1) +λ +� +B(∂q)B(∂¯q) +Mg2,n2 +� +κ=1 +B +� +V (2) +κ +�n1−1 +� +i=1 +V (1) +i +n2−1 +� +j=1 +V (2) +j +� +Σg,n +. +We introduce the surface states Σn1 and Σn2 such that the BPZ inner product with the +209 + +new states V (1) +n1 +and V (2) +n2 +reproduce the Mg1,n1- and Mg2,n2-forms: +⟨Σn1|V (1) +n1 ⟩ := ωMg1,n1(V (1) +1 +, . . . , V (1) +n1 ) = (2πi)−Mc +g1,n1 +�Mg1,n1 +� +λ=1 +B +� +V (1) +λ +� n1−1 +� +i=1 +V (1) +i +� +Σg1,n1 +, +(14.6a) +⟨Σn2|V (2) +n2 ⟩ := ωMg2,n2(V (2) +1 +, . . . , V (2) +n2 ) = (2πi)−Mc +g2,n2 +�Mg1,n1 +� +λ=1 +B +� +V (2) +λ +� n2−1 +� +j=1 +V (2) +j +� +Σg2,n2 +. +(14.6b) +As described in Section 13.1.2, these states exist since the p-form are linear in each of the +external state and the BPZ inner-product is non-degenerate. Each of the surface states +corresponds to an operator +⟨Σn1| =⟨0| I ◦ Σn1(0), +⟨Σn2| =⟨0| I ◦ Σn2(0), +(14.7) +defined from (6.136). Then, the forms can be interpreted as 2-point functions on the complex +plane: +⟨Σn1|V (1) +n1 ⟩ = ⟨I ◦ Σn1(0)Vn1(0)⟩w(1) +n1 , +⟨Σn2|V (2) +n2 ⟩ = ⟨I ◦ Σn2(0)Vn2(0)⟩w(2) +n2 . +(14.8) +All the complexity of the amplitudes has been lumped into the definitions of the surface +states which contain information about the surface moduli (including the ghost insertions) +and about the n1 − 1 remaining states (including the local coordinate systems). The local +coordinates around V (1) +n1 +and V (2) +n2 +are denoted respectively as w(1) +n1 and w(2) +n2 . Correspond- +ingly, the surface operators are inserted in the local coordinates w1 and w2 which are related +to w(1) +n1 and w(2) +n2 through the inversion: +w1 = I +� +w(1) +n1 +� +, +w2 = I +� +w(2) +n2 +� +. +(14.9) +In order to rewrite (14.5) in terms of Σ1 and Σ2, it is first necessary to express all +operators in one coordinate system, for example w(1) +n1 . Hence, we need to find its relation to +w2. Using the plumbing fixture (14.1), the relation between w(1) +n1 and w2 is: +w(1) +n1 = +q +w(2) +n2 += +q +I(w2) = qw2 := f(w2). +(14.10) +Then, the form (14.5) becomes +ωMg,n = +1 +2πi ⟨I ◦ Σn1(0)BqB¯q f ◦ Σn2(0)⟩w(1) +n1 = +1 +2πi ⟨Σn1| BqB¯q qL0 ¯q +¯L0 |Σ2⟩ , +(14.11) +using that Σ2 has a well-defined scaling dimension. The factor of 2πi arises by comparing the +contribution from Σn1 and Σn2 with the factor in (14.5). The expression can be simplified +by using the relation +⟨Σn1| BqB¯q |V (1) +n1 ⟩ = 1 +q¯q ⟨Σn1| b0¯b0 |V (1) +n1 ⟩ +(14.12) +using the expression (14.4) for Bq and the state–operator correspondence: +BqV (1) +n1 (z, ¯z) = 1 +q +� +Cq +dw(1) +n1 b +� +w(1) +n1 +� +w(1) +n1 V (1) +n1 (z, ¯z) −→ 1 +q b0 |V (1) +n1 ⟩ . +(14.13) +210 + +Ultimately, the form (14.5) reads +ωMg,n = +1 +2πi +1 +q¯q ⟨Σn1| b0¯b0 qL0 ¯q +¯L0 |Σn2⟩ . +(14.14) +It is important to remember that the plumbing fixture describes only a patch of the +moduli space, and the form defined in this way is valid only locally. As a consequence, the +integration over all moduli of Mg1,n1#Mg2,n2 does not describe Mg,n, but only a part of +it (Section 12.3.3). Every degeneration limit with a different puncture distribution in two +different groups contributes to a different part of the amplitude. +We denote the contribution to the total amplitude (14.2) from the region of the moduli +space connected to this degeneration limit as: +Fg,n +� +V (1) +i +|V (2) +j +� +:= +1 +2πi +� +Mg1,n1 +� +λ=1 +dt(1) +λ +Mg2,n2 +� +κ=1 +dt(2) +κ ∧ dq +q ∧ d¯q +¯q ⟨Σn1| b0¯b0 qL0 ¯q +¯L0 |Σn2⟩ . (14.15) +To proceed, we introduce a basis {φα(k)} of eigenstates of L0 and ¯L0, where kµ is the D- +dimensional momentum and α denotes the remaining quantum number. Then, introducing +twice the resolution of the identity (11.36) gives: +Fg,n +� +V (1) +i +|V (2) +j +� += +1 +2πi +� +dDk +(2π)D +dDk′ +(2π)D (−1)|φα| +× +� dq +q ∧ d¯q +¯q ⟨φα(k)c| b0¯b0 qL0 ¯q +¯L0 |φβ(k′)c⟩ +(14.16) +× +� +Mg1,n1 +� +λ=1 +dt(1) +λ ⟨Σn1|φα(k)⟩ +� +Mg2,n2 +� +κ=1 +dt(2) +κ ⟨φβ(k′)|Σn2⟩ +(with implicit sums over α and β). In the last line, one recognizes the expressions of the +g1-loop n1-point amplitude with external states {V (1) +1 +, . . . , V (1) +n1−1, φα} and of the g2-loop +and n2-point amplitudes with external states {V (2) +1 +, . . . , V (2) +n2−1, φβ}: +Ag1,n1 +� +V (1) +1 +, . . . , V (1) +n1−1, φα(k) +� += +� +Sg1,n1 +ωMg1,n1 +� +V (1) +1 +, . . . , V (1) +n1−1, φα(k) +� += +� +Sg1,n1 +Mg1,n1 +� +λ=1 +dt(1) +λ ⟨Σn1|φα(k)⟩ , +(14.17a) +Ag2,n2 +� +V (2) +1 +, . . . , V (2) +n2−1, φβ(k′) +� += +� +Sg2,n2 +ωMg2,n2 +� +V (2) +1 +, . . . , V (2) +n2−2, φβ(k′) +� += +� +Sg2,n2 +Mg2,n2 +� +λ=1 +dt(2) +λ ⟨Σn2|φβ(k′)⟩ . +(14.17b) +The property (B.27) has been used to reverse the order of the BPZ product for the second +Riemann surface, and this cancels the factor (−1)|φα|. +Defining the second line of (14.16) as +∆αβ(k, k′) := ∆ +� +φα(k)c, φβ(k′)c� +:= +1 +2πi +� dq +q ∧ d¯q +¯q ⟨φα(k)c| b0¯b0 qL0 ¯q +¯L0 |φβ(k′)c⟩ , (14.18) +one has: +Fg,n +� +V (1) +i +|V (2) +j +� += +� +dDk +(2π)D +dDk′ +(2π)D Ag1,n1 +� +V (1) +1 +, . . . , V (1) +n1−1, φα(k) +� +∆αβ(k, k′) +× Ag2,n2 +� +V (2) +1 +, . . . , V (2) +n2−1, φβ(k′) +� +. +(14.19) +211 + +We recover the expressions from Section 11.1.2, but for a more general amplitude. +We +had found that ∆ corresponds to the propagator: its properties are studied further in +Section 14.2.2. +Hence, the object (14.16) corresponds to the product of two amplitudes +connected by a propagator (Figure 14.2). +There are several points to mention about this amplitude: +• We will find that the propagator depends only on one momentum because ⟨k|k′⟩ ∼ +δ(D)(k + k′), which removes one of the integral. Then, both amplitudes Ag1,n1 and +Ag2,n2 contain a delta function for the momenta: +Ag1,n1 ∼ δ(D)� +k(1) +1 +· · ·+k(1) +n1−1+k +� +, +Ag2,n2 ∼ δ(D)� +k(2) +1 +· · ·+k(2) +n2−1+k′� +. (14.20) +As a consequence, the second momentum integral can be performed and yields a delta +function: +Fg,n ∼ δ(D)� +k(1) +1 ++ · · · + k(1) +n1−1 + k(2) +1 ++ · · · + k(2) +n2−1 +� +. +(14.21) +Hence, the momentum flowing in the internal line is fixed and this ensures the overall +momentum conservation as expected. +• The ghost numbers of the states φα and φβ are also fixed (in terms of the external +states). Indeed, because of the ghost number anomaly, the amplitudes on Mg1,n1 and +Mg2,n2 are non-vanishing only if the ghost numbers of these states satisfy: +Ngh(φα) = 2n1 − +n1−1 +� +i=1 +Ngh +� +V (1) +i +� +, +Ngh(φβ) = 2n2 − +n2−1 +� +j=1 +Ngh +� +V (2) +j +� +. +(14.22) +The non-vanishing of Fg,n also gives another relation: +Ngh(φα) + Ngh(φβ) = 4. +(14.23) +In particular, if the external states are on-shell with Ngh = 2, we find: +Ngh(φα) = Ngh(φβ) = 2. +(14.24) +As indicated in Chapter 10, such states are appropriate at the classical level since they +do not contain spacetime ghosts. +• The sum over α and β is over an infinite number of states and could diverge. In fact, +the sum can be made convergent by tuning the stub parameter (Section 14.2.4). +Properties of Feynman graphs and amplitudes in the momentum space will be discussed +further in Chapter 18. +14.1.2 +Non-separating case +Next, we consider the non-separating plumbing fixture (Section 12.3.2). The computations +are almost identical to the separating case, thus we outline only the general steps. +Part of the moduli space Mg,n is covered by #Mg1,n1, with g = g1 + 1 and n = n1 − 2. +The local coordinates are denoted as wi for i = 1, . . . , n1 and the plumbing fixture reads: +wn1−1wn1 = q. +(14.25) +The g-loop n-point amplitude with external states {V (1) +1 +, . . . , V (1) +n1−2} is denoted as: +Ag,n = +� +Sg,n +ωg,n +Mg,n +� +V (1) +1 +, . . . , V (1) +n1−2 +� +. +(14.26) +212 + +Figure 14.2: Factorization of the amplitude into two sub-amplitudes connected by a propag- +ator (dashed line). +When the n1 − 2 punctures and g1 = g − 1 holes move lose to each other, the form can +be written as: +ωMg,n +� +V (1) +1 +, . . . , V (1) +Mg1,n1, ∂q, ∂¯q +� += (2πi)−Mc +g,n +�Mg1,n1 +� +λ=1 +B +� +V (1) +λ +� +B(∂q)B(∂¯q) +n1−2 +� +i=1 +V (1) +i +� +Σg,n +. +(14.27) +To proceed, one needs to introduce the surface state Σn1−1,n1: +⟨Σn1−1,n1|V (1) +n1−1 ⊗ V (1) +n1 ⟩ := ωMg1,n1(V (1) +1 +, . . . , V (1) +n1 ). +(14.28) +Following the same step as in the previous section leads to: +Fg,n +� +V (1) +i +| +� += +� +dDk +(2π)D +dDk′ +(2π)D Ag1,n1 +� +V (1) +1 +, . . . , V (1) +n1−2, φα(k), φβ(k′) +� +∆αβ(k, k′), +(14.29) +where the propagator is given in (14.18). This is equivalent to an amplitude for which two +external legs are glued together with a propagator, giving a loop (Figure 14.3). +Since both states φα and φβ are inserted on the same surface, their ghost numbers are +not fixed, even if the external states are physical. The non-vanishing of Fg,n only leads to +the constraint: +Ngh(φα) + Ngh(φβ) = 2n1 − +n1−2 +� +i=1 +Ngh +� +V (1) +i +� += 4. +(14.30) +As a consequence, loop diagrams force to introduce states of every ghost number. Internal +states with Ngh ̸= 2 correspond to spacetime ghosts. +Since the propagator contains a delta function δ(D)(k − k′), the integral over k′ can be +removed by setting k′ = −k. However, the integral over k remains since +Ag1,n1 +� +V (1) +1 +, . . . , V (1) +n1−2, φα(k), φβ(−k) +� +∼ δ(D)� +k(1) +1 ++ · · · + k(1) +n1−2 +� +. +(14.31) +Hence, the loop momentum k is not fixed, as expected in QFT. +Remark 14.1 Not all values of the moduli associated to the holes can be associated to loops +in Feynman diagrams. Only the values close to the degeneration limit can be interpreted in +this way, the other being just standard (quantum) vertices. +14.2 +Feynman diagrams and Feynman rules +In the standard QFT approach, Feynman graphs compute Green functions, and scattering +amplitudes are obtained by amputating the external propagators through the LSZ prescrip- +tion. For connected tree-level processes, this requires n ≥ 3 (corresponding to χ0,n < 0). +213 + +Figure 14.3: Factorization of the amplitude into two sub-amplitudes connected by a propag- +ator (dashed line). The propagator connects two punctures of the same surface, which is +equivalent to a loop. +Given a theory, there is a minimal set of Feynman diagrams – the Feynman rules – from +which every other diagram can be constructed. These rules include the definitions of the +fundamental vertices – the fundamental interactions – and of the propagator – how states +propagate between two interactions (or, how to glue vertices together). In this section, we +describe these different elements. +14.2.1 +Feynman graphs +The amplitude factorization described in Section 14.1 gives a natural separation of amp- +litudes into several contributions. Considering all the possible degeneration limits lead to +a set of diagrams with amplitudes of lower order connected by propagators (Figure 14.2, +Figure 14.3). This corresponds exactly to the idea behind Feynman graphs. Then, the goal +is to find the Feynman rules of the theory: since the propagator has already been identified +(further studied in Section 14.2.2), it is sufficient to find the interaction vertices. +Let’s make this more precise by considering an amplitude Ag,n(V1, . . . , Vn). The index +of an amplitude is defined to be the index (12.50) of the corresponding Riemann surfaces +r(Ag,n) := r(Σg,n) = 3g + n − 2. +(14.32) +Contributions to an amplitude with a given r(Ag,n) can be described in terms of amp- +litudes Ag′,n′ with r(Ag′,n′) < r(Ag,n). But, the moduli space Mg,n cannot (generically) be +completely covered with the plumbing fixture of lower-dimensional moduli spaces, i.e. with +r(Mg′,n′) < r(Mg,n) (Section 12.3.3). Then, the same must be true for the amplitudes, +such that Ag,n cannot be uniquely expressed in terms of amplitudes Ag′,n′. +The g-loop n-point fundamental vertex is defined by: +Vg,n(V1, . . . , Vn) := +:= +� +Rg,n +ωg,n +Mg,n(V1, . . . , Vn), +(14.33) +The form defined in (13.16) is integrated over a sub-section Rg,n ⊂ Sg,n of ˆPg,n. Its pro- +jection on the base is the region Vg,n ⊂ Mg,n which cannot be described by the plumbing +214 + +fixture, see (12.42b). In general, we will keep the choice of local coordinates implicit and +always write Vg,n to avoid surcharging the notations. +It corresponds to the remaining contribution of the amplitude once all graphs containing +propagators have been taken into account: += +� +0≤h≤g +0≤m 0. As argued in the introduction, divergences +for L+ +0 ≤ 0 are either non-physical or IR divergences which can be cured by renormalization. +For this reason, we take the RHS as a definition of the integral, which would be the correct +result if one starts with a field theory action instead of a first-quantized formalism. +In this case, the propagator becomes +∆ = b+ +0 +L+ +0 +b− +0 δL− +0 ,0. +(14.41) +This is the standard expression for the propagator. For completeness, the form in terms of +the holomorphic and anti-holomorphic components is: +∆ = −2b0¯b0 +1 +L0 + ¯L0 +δL0,¯L0. +(14.42) +The delta function restricts the amplitude to states satisfying the level-matching condition, +that is, annihilated by L− +0 . +Considering a basis {φα(k)} of eigenstates of both L0 and ¯L0: +L+ +0 |φα(k)⟩ = α′ +2 (k2 + m2 +α) |φα(k)⟩ , +L− +0 |φα(k)⟩ = 0 +(14.43) +leads to the following momentum-space kernel for the propagator: +∆αβ(k, k′) :=⟨φα(k)c| ∆ |φβ(k′)c⟩ := (2π)Dδ(D)(k + k′) ∆αβ(k), +(14.44a) +∆αβ(k) := Mαβ(k) +k2 + m2α +, +Mαβ(k) := 2 +α′ ⟨φc +α(k)| b+ +0 b− +0 |φc +β(−k)⟩ , +(14.44b) +with Mαβ a finite-dimensional matrix giving the overlap of states of identical masses (because +the number of states at a given level is finite). +For the propagator to be well-defined, it must be invertible (in particular, to define a +kinetic term). The propagator (14.41) is non-vanishing if the states it acts on satisfy: +b+ +0 |φc +α⟩ ̸= 0, +b− +0 |φc +α⟩ ̸= 0. +(14.45) +216 + +Necessary and sufficient conditions for this to be true are +c+ +0 |φc +α⟩ = 0, +c− +0 |φc +α⟩ = 0. +(14.46) +Indeed, decomposing the state on the ghost zero-modes +|φc +α⟩ = |φ1⟩ + b± +0 |φ2⟩ , +c± +0 |φ1⟩ = c± +0 |φ2⟩ = 0 +(14.47) +gives +c± +0 |φc +α⟩ = 0 +=⇒ +|φ2⟩ = 0, +(14.48) +and one has correctly b± +0 |φ1⟩ ̸= 0. +These conditions are given for the dual states: translating them on the normal states +reverses the roles of b0 and c0. Hence, the states must satisfy the conditions: +b+ +0 |φα⟩ = 0, +b− +0 |φα⟩ = 0. +(14.49) +The second condition is satisfied automatically because the Hilbert space is H− when work- +ing with ˆPg,n (Section 13.3.1). However, the first condition further restricts the states which +propagate in internal lines. This leads to postulate that the external states should also be +taken to satisfy this condition +b+ +0 |Vi⟩ = 0, +(14.50) +since external states are usually a subset of the internal states. This provides another mo- +tivation of the statement in Section 3.2.2 that scattering amplitudes for the states not anni- +hilated by b+ +0 must be trivial. A field interpretation of this condition is given in Chapters 10 +and 15. +Under these constraints on the states, the propagator can be inverted: +∆−1 = c+ +0 c− +0 L+ +0 δL− +0 ,0. +(14.51) +14.2.3 +Fundamental vertices +The vertices (14.33) can be constructed recursively assuming that all amplitudes are known. +The starting point is the tree-level cubic amplitude A0,3: since it does not contain any +internal propagator, it is equal to the fundamental vertex V0,3. +The fist thing to extract from the recursion relations are the background independent +data. This amounts to find local coordinates and a characterization of the subspaces Vg,n ⊂ +Mg,n, starting with P0,3 and iterating. +In the rest of this section, we show how this works schematically. +Recursive definition: tree-level vertices +The description of tree-level amplitudes A0,n is the simplest since only the separating plumb- +ing fixture is used and Feynman graphs are trees. The possible factorizations of the amp- +litude correspond basically to all the partitions of the set {Vi} into subsets. +Tree-level cubic vertex +Since M0,3 = 0, the moduli space of the 3-punctured sphere +Σ0,3 reduces to a point, and so does the section S0,3 of P0,3 (Figure 14.4a): +V0,3(V1, V2, V3) := A0,3(V1, V2, V3) = ω0,3 +0 (V1, V2, V3). +(14.52) +The corresponding graph is indicated in Figure 14.4b. +217 + +(a) A section S0,3 over P0,3 reduces to +a point. +(b) Fundamental cubic vertex. +Figure 14.4: Section of P0,3 and cubic vertex. +Tree-level quartic vertex +Part of the contributions to the 4-point amplitude A0,4 with +external states Vi (i = 1, . . . , 4) comes from gluing two cubic vertices. Because there are four +external states, there are three different partitions 2|2 which are described in Figure 14.5 +(see also Figure 12.7). The sum of these three diagrams does not reproduce A0,4: the moduli +space M0,4 is not completely covered by the three amplitudes. Equivalently, the projection +of the section over P0,4 does not cover all of M0,4. The missing contribution is defined by +the quartic vertex (Figure 14.7) +V0,4(V1, V2, V3, V4) := +� +R0,4 +ω0,4 +2 (V1, . . . , V4), +(14.53) +and the corresponding section is denoted by R0,4 (Figure 14.6). Denoting by F(s,t,u) +0,4 +the +graphs 14.5 in the s-, t- and u-channels, one has the relation +A0,4 = F(s) +0,4 + F(t) +0,4 + F(u) +0,4 + V0,4. +(14.54) +Tree-level quintic vertex +The amplitude A0,5 can be factorized in a greater number +of channels, the two types being 2|2|1 and 2|3. +The possible Feynman graphs are built +either from three cubic vertices and two propagators (Figure 14.8a and permutations), or +from one cubic and one quartic vertices together with one propagator (Figure 14.8b and +permutations). The remaining contribution is the fundamental vertex (Figure 14.8c): +V0,5(V1, . . . , V5) := +� +R0,5 +ω0,5 +4 (V1, . . . , V5). +(14.55) +The construction to higher-order follows exactly this scheme. +Recursive definition: general vertices +Next, one needs to consider Feynman diagrams with loops. The first amplitude which can +be considered is the one-loop tadpole A1,1(V1). +The factorization region corresponds to +the graph obtained by gluing two legs of the cubic vertex (Section 14.2.3). The remaining +contribution is the fundamental tadpole vertex V1,1(V1) (Section 14.2.3) – note the index +g = 1 on the vertex, indicating that it is a 1-loop effect. +Next, the 1-loop 2-point amplitude can be obtained using the cubic and quartic tree-level +vertices V0,3 and V0,4, but also the one-loop tadpole V1,1. Iterating, the number of loops +218 + +(a) s-channel. +(b) t-channel. +(c) u-channel. +Figure 14.5: Factorization of the quartic amplitude A0,4 in the s-, t- and u-channels. +Figure 14.6: A section S0,4 over P0,4, the contribution from the s-, t- and u-channels (Fig- +ure 14.5) are indicated by the corresponding indices. The fundamental vertex is defined by +the section V0,4. +219 + +Figure 14.7: Fundamental quartic vertex. +(a) Factorization 12|3|45. +(b) Factorization 12|345. +(c) Fundamental vertex. +Figure 14.8: Factorization of the amplitude G0,5 in channels and fundamental quintic vertex. +Only the cases where V1 and V2 factorize on one side is indicated, the other cases follow by +permutations of the external states. +220 + +can be increased either by gluing together two external legs of a graph, or by gluing two +different graphs with loops together. +For g ≥ 2, the recursion implies the existence of vertices with no external states Vg,0: +they should be interpreted as loop corrections to the vacuum energy density. +It is important to realize that, in this language, a handle in the Riemann surface is +not necessarily mapped to a loop in the Feynman graph: only handles described by the +region Fg,n = Mg,n − Vg,n do. The higher-order vertices – corresponding to surfaces with +small handles only and described by Vg,n – should be regarded as quantum fundamental +interactions. +In Chapter 15, it will be explained that they really correspond to (finite) +counter-terms: the measure is not invariant under the gauge symmetry of the theory and +these terms must be introduced to restore it. +(a) Internal loop. +(b) Fundamental vertex. +Figure 14.9: Factorization of the amplitude G1,1 and fundamental tadpole at 1-loop. +Other vertices +The definition given at the end of (14.2.1) suggests to introduce additional vertices. The +previous recursive definition gives only vertices with χg,n = 2 − 2g − n < 0, but, in fact, it +makes sense to consider the additional cases: g = 0 and n = 0, 1, 2, and g = 1, n = 0. +The definition of the vertices as amputated Green function without internal propagators +provides a hint for the tree-level quadratic vertex V0,2. We define the latter as the amputated +tree-level 2-point Green function: +V0,2 := ∆−1∆∆−1 = ∆−1. +(14.56) +Hence, we have +V0,2(V1, V2) :=⟨V1| c+ +0 c− +0 L+ +0 δL− +0 ,0 |V2⟩ . +(14.57) +Note that V0,2 is not the 2-point scattering amplitude. +We denote the tree-level 1-point and 0-point vertices as V0,1(V1) and V0,0. The first can +be interpreted as a classical source in the action, while the second is a classical vacuum +energy. They are set to zero in most applications and can be safely ignored. However, they +appear when formulating the theory on a background which does not solve the equation of +motion [263]. +Finally, the 1-loop vacuum energy V1,0 can also be defined as the partition function of +the worldsheet CFT integrated over the torus modulus. +This allows to define the vertices Vg,n for all g, n ∈ N. We define the sum of all loop +contributions for a fixed n as: +Vn(V1, . . . , Vn) := +� +g≥0 +(ℏg2 +s)g Vg,n(V1, . . . , Vn). +(14.58) +221 + +14.2.4 +Stubs +In Section 12.3.4, we have indicated that the plumbing fixture can be modified by adding +stubs or, equivalently, by rescaling the local coordinates. This amounts to introduce a cut-off +(12.53) on the variable s such that +q = e−s+iθ, +s ∈ [s0, ∞), +θ ∈ [0, 2π). +(14.59) +instead of (14.37). In this case, the s-integral in the propagator (14.36) is modified to +� ∞ +s0 +ds e−sL+ +0 = e−s0L+ +0 +L+ +0 +. +(14.60) +This leads to a new expression for the propagator: +∆(s0) = b+ +0 +e−s0L+ +0 +L+ +0 +b− +0 δL− +0 ,0. +(14.61) +In momentum space, this reads +∆αβ(k) := e− α′s0 +2 +(k2+m2 +α) +k2 + m2α +Mαβ(k). +(14.62) +It is more convenient to work with the canonical propagator (14.41). This can be achieved +by absorbing e− s0 +2 L+ +0 in the interaction vertex: a n-point interaction will get n such factors.1 +Since s0 changes the local coordinates, this means that it also changes the region Vg,n +(Figure 12.10). The freedom in the choice of s0 translates into a freedom to choose which part +of the amplitude is described by propagator graphs Fg,n(s0), and which part is described by +a fundamental vertex Vg,n(s0). The amplitude Ag,n is independent of s0 since it is described +in terms of the complete moduli space Mg,n. This also means that the parameter s0 must +disappear when summing over the contributions from Vg,n(s0) and Fg,n(s0). This indicates +that the value of s0 is not relevant, even off-shell: it can be taken to any convenient value. +The possibility of adding stubs solves the problem that the sum over all states could +diverge (see Section 14.1.1). Indeed, the expression (14.62) in momentum space shows that +the propagator includes an exponential suppression for very massive particle propagating as +intermediate states. Since the mass of a particle increases with the level, this shows that the +sum converges for a sufficiently large value of s0 thanks to the factor e−α′s0m2. A second +interesting aspect is the exponential momentum suppression e−α′s0k2: this is responsible for +the nice UV behaviour of string theory. Since the value of s0 is not physical, this means +that all Feynman graphs must share these properties. These two points will be made more +precise in Chapter 18. +14.2.5 +1PI vertices +We can follow the same procedure as before, but considering only the separating plumb- +ing fixture. In this case, the Feynman diagrams are all 1PR (1-particle reducible): if the +propagator line is cut, then the graphs split in two disconnected components. The region +of the moduli space covered by these graphs is written as F1PR +g,n +(12.43a). The complement +1To make this identification precise for vertices involving external states, one has to consider the non- +amputated Green functions. +222 + +defines the 1PI region V1PI +g,n (12.43b). Then, the 1PI g-loop n-point fundamental vertices are +defined as: +V1PI +g,n (V1, . . . , Vn) := +:= +� +R1PI +g,n +ωg,n +Mg,n(V1, . . . , Vn), +(14.63) +where R1PI +g,n is a section of Pg,n which projection on the base is V1PI +g,n . +14.3 +Properties of fundamental vertices +14.3.1 +String product +Following the definition of surfaces states (Section 13.1.2), the vertex state is defined as: +⟨Vg,n| ⊗i Vi⟩ := Vg,n(⊗iVi). +(14.64) +The vertex is a map Vg,n : H⊗n → C where C ≃ H⊗0. We will find very useful to +introduce the string products ℓg,n : H⊗n → H through the closed string inner product: +Vg,n+1(V0, V1, . . . , Vn) :=⟨V0| c− +0 |ℓg,n(V1, . . . , Vn)⟩ . +(14.65) +An alternative notation is: +ℓg,n(V1, . . . , Vn) := [V1, . . . , Vn]g +(14.66) +The advantage of the second notation is to show that the products with n ≥ 3 are direct +generalization of the 2-product, which is very similar to a super-Lie bracket. These products +play a central role in SFT – in fact, the description of SFT is more natural using the ℓg,n +rather than the Vg,n. +Note that the products with n = 0 are maps C → H, which means that they correspond +to a particular fixed state. +ℓg,0 := [·]g ∈ H. +(14.67) +The ghost number of the product (14.65) is +Ngh +� +ℓg,n(V1, . . . , Vn) +� += 3 − 2n + +n +� +i=1 +Ngh(Vi) = 3 + +n +� +i=1 +� +Ngh(Vi) − 2 +� +, +(14.68) +and it is independent of the genus g. As a consequence, the parity of the product is +|ℓg,n(V1, . . . , Vn)| = 1 + +n +� +i=1 +|Vi| +mod 2, +(14.69) +and the string product itself is always odd. +The vertices satisfy the following identity for g ≥ 0 and n ≥ 1 [262, pp. 41–42] +0 = +� +g1,g2≥0 +g1+g2=g +� +n1,n2≥0 +n1+n2=n +n! +n1! n2!Vg1,n1+1 +� +Ψn1, ℓg2,n2(Ψn2) +� ++ (−1)|φs|Vg−1,n+2 +� +φs, b− +0 φc +s, Ψn� +. +(14.70) +The last term is absent for g = 0. It is a consequence of the definition of the vertices as the +missing region from gluing lower-order vertices. +223 + +14.3.2 +Feynman graph interpretation +The vertices must satisfy a certain number of conditions to be interpreted as Feynman +diagrams. The first is that they must be symmetric under permutations of the states. Not +every choice of local coordinates satisfies this requirement: this can be solved by defining +the vertex over a generalized section. In this case, the vertex is defined as the average of +the integrals over N sections S(a) +g,n of Pg,n: +Vg,n(V1, . . . , Vn) = 1 +N +N +� +a=1 +� +R(a) +g,n +ωg,n +Mg,n(V1, . . . , Vn). +(14.71) +Example 14.2 – 3-point vertex +The cubic vertex must be symmetric under permutations +V0,3(V1, V2, V3) = V0,3(V3, V1, V2) + · · · +(14.72) +Taking the vertex to be given by a section S0,3 with local coordinates fi +V0,3(V1, V2, V3) = ω0,3 +0 (V1, V2, V3)|S0,3 = ⟨f1 ◦ V1(0)f2 ◦ V2(0)f3 ◦ V3(0)⟩, +(14.73) +one finds that a permutation looks different +V0,3(V3, V1, V2) = ⟨f1 ◦ V3(0)f2 ◦ V1(0)f3 ◦ V2(0)⟩ ̸= V0,3(V1, V2, V3), +(14.74) +unless the local coordinates satisfy special properties (remember that the local co- +ordinates are specified by the vertex state V and not by the external states Vi, so a +permutation of them does not permute the local maps). Obviously, both amplitudes +agree on-shell since the dependence in the local coordinates cancel (equivalently one +can rotate the punctures using SL(2, C)). +Writing zi = fi(0), there is a SL(2, C) transformation g(z) such that +g(z1) = z2, +g(z2) = z3, +g(z3) = z1 +(14.75) +such that +V0,3(V3, V1, V2) = ⟨g ◦ f1 ◦ V3(0)g ◦ f2 ◦ V1(0)g ◦ f3 ◦ V2(0)⟩. +(14.76) +While the state Vi is correctly inserted at the puncture zi in this expression, this is not +sufficient to guarantee the equality of the amplitudes. Indeed the fibre is defined by the +complete functions fi(w) and not only by their values at w = 0. For this reason the +amplitudes can be equal only if +g ◦ f1 = f2, +g ◦ f2 = f3, +g ◦ f3 = f1. +(14.77) +This provides constraints on the functions fi, but it is often not possible to solve them. +If the constraints cannot be solved, then one must introduce a general section. In +this case a generalized section will be made of 6 sections S(a) (a = 1, . . . , 6) because +there are 6 permutations. Then the amplitude reads +V0,3(V1, V2, V3) = 1 +6 +6 +� +a=1 +ω0,3 +0 (V1, V2, V3)|S(a) +0,3 . +(14.78) +224 + +Figure 14.10: A generalized section {S(a) +0,3} (a = 1, . . . , 6) of P0,3 for the 3-point vertex. This +is to be compared with Figure 14.4a. +When computing the Feynman graphs by gluing lower-dimensional amplitudes, it is +possible that parts of the section overlap, meaning that several graphs cover the same part +of the moduli space. In this case, the fundamental vertex should be defined as a negative +contribution in the overlap region. This procedure is perfectly well-defined since all graphs +are finite and there is no ambiguity. In practice, it is always simpler to work with non- +overlapping sections (i.e. a single covering of the moduli space). A simple way to prevent +overlaps is to tune the stub parameter s0 to a large value. +By construction, the integral over Vg,n should be finite. If this is not the case, it means +that the propagator graphs also diverge and that the parametrization is not good. This can +also be solved by considering a sufficiently large value of the stub parameter s0. +14.4 +Suggested readings +• Plumbing fixture and amplitude factorization [193, sec. 9.3, 9.4, 256, sec. 6]. +225 + +Chapter 15 +Closed string field theory +Abstract +We bring together the elements from the previous chapters in order to write +the closed string field action. We first study the gauge fixed theory before reintroducing the +gauge invariance. We then prove that the action satisfies the BV master equation meaning +that closed SFT is completely consistent at the quantum level. Finally, we describe the 1PI +effective action. +15.1 +Closed string field expansion +In Chapters 11, 13 and 14, constraints on the external and internal states were found to +be necessary. But, to provide another perspective and decouples the properties of the field +from the ones of the state, we assume that the string field does not obey any constraint. +They will be derived later in order to reproduce the scattering amplitudes from the action +and to make the latter well-defined. +The string field is expanded on a basis {φr} of the CFT Hilbert space H (see Section 11.2 +for more details) +|Ψ⟩ = +� +r +ψr |φr⟩ . +(15.1) +Using the decomposition (11.40) of the Hilbert space according to the ghost zero-modes, the +string field can also be expanded as +|Ψ⟩ = +� +r +� +ψ↓↓,r |φ↓↓,r⟩ + ψ↓↑,r |φ↓↑,r⟩ + ψ↑↓,r |φ↑↓,r⟩ + ψ↑↑,r |φ↑↑,r⟩ +� +, +(15.2) +where we recall that the basis states satisfy +b0 |φ↓↓,r⟩ = ¯b0 |φ↓↓,r⟩ = 0, +b0 |φ↓↑,r⟩ = ¯c0 |φ↓↑,r⟩ = 0, +c0 |φ↑↓,r⟩ = ¯b0 |φ↑↓,r⟩ = 0, +c0 |φ↑↑,r⟩ = ¯c0 |φ↑↑,r⟩ = 0. +(15.3) +We recall the definition of the dual basis {φc +r} through the BPZ inner product +⟨φc +r|φs⟩ = δrs. +(15.4) +In terms of the ghost decomposition, the components of the dual states satisfy: +⟨φc +↓↓,r| c0 =⟨φc +↓↓,r| ¯c0 = 0, +⟨φc +↓↑,r| c0 =⟨φc +↓↑,r|¯b0 = 0, +⟨φc +↑↓,r| b0 =⟨φc +↑↓,r| ¯c0 = 0, +⟨φc +↑↑,r| b0 =⟨φc +↑↑,r|¯b0 = 0, +⟨φc +x,r|φy,s⟩ = δxyδrs, +(15.5) +226 + +where x, y =↓↓, ↑↓, ↓↑, ↑↑. The spacetime ghost number of the fields ψr is defined by +G(ψr) = 2 − nr. +(15.6) +Remember that the ghost number of the basis states are denoted by +nr = Ngh(φr), +nc +r = Ngh(φc +r) = 6 − nr. +(15.7) +15.2 +Gauge fixed theory +Having built the kinetic term (Chapter 9), one needs to construct the interactions. For +the same reason – our ignorance of SFT first principles – that forced us to start with the +free equation of motion to derive the quadratic action (Chapter 10), we also need to infer +the interactions from the scattering amplitudes. Preparing the stage for this analysis was +the goal of Chapter 14, where we introduced the factorization of amplitudes to derive the +fundamental interactions. +Scattering amplitudes are expressed in terms of gauge fixed states since only them are +physical. This allows to give an alternative derivation of the kinetic term by defining it as +the inverse of the propagator, which is well-defined for gauge fixed states.1 The price to +pay by constructing interactions in this way is that the SFT action itself is gauge fixed. To +undercover its deeper structure it is necessary to release the gauge fixing condition. In view +of the analysis of the quadratic action in Chapter 10, we can expect that the BV formalism +is required. Another possibility is to consider directly the 1PI action. +In this section, we first derive the kinetic term by inverting the propagator. For this to +be possible, the string field must obey some constraints: we will find that they correspond +to the level-matching and Siegel gauge conditions. Then, we introduce the interactions into +the action. +15.2.1 +Kinetic term and propagator +In Chapter 14, it was found that the propagator reads (14.41): +∆ = b+ +0 b− +0 +1 +L+ +0 +δL− +0 ,0, +∆rs =⟨φc +r| b+ +0 b− +0 +1 +L+ +0 +δL− +0 ,0 |φc +s⟩ . +(15.8) +The most natural guess for the kinetic term is +S0,2 = 1 +2 ⟨Ψ| K |Ψ⟩ = 1 +2 ψrKrsψs +(15.9) +where +K = c− +0 c+ +0 L+ +0 δL− +0 ,0 +Krs =⟨φr| c− +0 c+ +0 L+ +0 δL− +0 ,0 |φs⟩ . +(15.10) +Indeed, it looks like K∆ = 1 using the identities c± +0 b± +0 ∼ 1 and it matches (10.115). In +terms of the holomorphic and anti-holomorphic modes, we have +K = 1 +2 c0¯c0L+ +0 δL− +0 ,0. +(15.11) +But, when writing c± +0 b± +0 ∼ 1, the second part of the anti-commutator {b± +0 , c± +0 } = 1 is +missing. The relation c± +0 b± +0 ∼ 1 is correct only when acting on basis dual states annihilated +1This step is not necessary because the propagator corresponding to the plumbing fixture (Section 14.2.2) +matches the one found in Section 10.5 by considering the simplest gauge fixing. However, this would have +been necessary if the factorization had given another propagator, or if the structure of the theory was more +complicated, for example for the superstring. +227 + +by c± +0 . +The problem stems from the fact that Ψ is not yet subject to any constraint. +Moreover, some of the string field components will not appear in the expression since they +are annihilated by the ghost zero-mode. As a consequence, the kinetic operator in (15.10) +(or equivalently the propagator) is not invertible in the Hilbert space H because its kernel +is not empty: +ker K|H ̸= ∅. +(15.12) +This can be seen by writing φr as a 4-vector and Krs as a 4 × 4-matrix: +Krs = 1 +2 +� +� +� +� +⟨φ↓↓,r| +⟨φ↓↑,r| +⟨φ↑↓,r| +⟨φ↑↑,r| +� +� +� +� +t � +� +� +� +c0¯c0L+ +0 +0 +0 +0 +0 +0 +0 +0 +0 +0 +0 +0 +0 +0 +0 +0 +� +� +� +� +� +� +� +� +|φ↓↓,s⟩ +|φ↓↑,s⟩ +|φ↑↓,s⟩ +|φ↑↑,s⟩ +� +� +� +� . +(15.13) +The matrix is mostly empty because the states φx,r with different x =↓↓, ↑↓, ↓↑, ↑↑ are +orthogonal (no non-diagonal terms) and the states with x ̸=↓↓ are annihilated by c0 or ¯c0. +The same consideration applies for the delta-function: if the field does not satisfy L− +0 = 0, +then the kinetic operator is non-invertible. +To summarize the string field must satisfy three conditions in order to have an invertible +kinetic term +L− +0 |Ψ⟩ = 0, +b− +0 |Ψ⟩ = 0, +b+ +0 |Ψ⟩ = 0. +(15.14) +This means that the string field is expanded on the H0 ∩ ker L− +0 Hilbert space: +|Ψ⟩ = +� +r +ψ↓↓,r |φ↓↓,r⟩ . +(15.15) +Ill-defined kinetic terms are expected in the presence of a gauge symmetry: this was already +discussed in Sections 10.1.4 and 10.5 for the free theory, and this will be discussed further +later in this chapter for the interacting case. +Computation +Let’s check that Krs is correctly the inverse of ∆rs when Ψ is restricted to H0: +Krs∆st =⟨φr| c− +0 c+ +0 L+ +0 δL− +0 ,0 |φs⟩⟨φc +s| b+ +0 b− +0 +1 +L+ +0 +δL− +0 ,0 |φc +t⟩ +=⟨φr| c− +0 c+ +0 L+ +0 δL− +0 ,0b+ +0 b− +0 +1 +L+ +0 +δL− +0 ,0 |φc +t⟩ +=⟨φr| {c− +0 , b− +0 }{c+ +0 , b+ +0 } |φc +t⟩ += ⟨φr|φc +t⟩ = δrt. +The second equality follows from the resolution of the identity (11.36): due to the zero- +mode insertions, the resolution of the identity collapses to a sum over the ↓↓ states +1 = +� +r +|φr⟩⟨φc +r| = +� +r +|φ↓↓,r⟩⟨φc +↓↓,r| . +(15.16) +The third equality uses that L+ +0 commutes with the ghost modes, that φr is annihilated +by b± +0 , and that (δL− +0 ,0)2 = δL− +0 ,0 = 1 on states with L− +0 = 0. +Finally, we find that the kinetic term matches the classical quadratic vertex V0,2 defined +in (14.56) such that +S0,2 = 1 +2 V0,2(Ψ2) = 1 +2 ⟨Ψ| c− +0 c+ +0 L+ +0 δL− +0 ,0 |Ψ⟩ . +(15.17) +228 + +15.2.2 +Interactions +The second step to build the action is to write the interaction terms from the Feynman rules. +Before proceeding to SFT, it is useful to remember how this works for a standard QFT. +Example 15.1 – Feynman rules for a scalar field +Consider a scalar field with a standard kinetic term and a n-point interaction: +S = +� +dDx +�1 +2 φ(x)(−∂2 + m2)φ(x) + λ +n! φ(x)n +� +. +(15.18) +First, one needs to find the physical states, which correspond to solutions of the lin- +earised equation of motion. In the current case, they are plane-waves (in momentum +representation): +φk(x) = eik·x. +(15.19) +Then, the vertex (in momentum representation) Vn(k1, . . . , kn) is found by replacing +in the interaction each occurrence of the field by a different state, and summing over +all the different contributions. Here, this means that one considers states φki(x) with +different momenta: +Vn(k1, . . . , kn) = λ +n! +� +dDx n! +n +� +i=1 +φki(x) = λ +� +dDx ei(k1+···+kn)x += λ(2π)D δ(D)(k1 + · · · + kn). +(15.20) +The factor n! comes from all the permutations of the n states in the monomial of order +n. Reversing the argument, one sees how to move from the vertex Vn(k1, . . . , kn) written +in terms of states to the interaction in the action in terms of the field. +Obviously, if the field has more states (for example if it has a spin or if it is in a +representation of a group), then one needs to consider all the different possibilities. The +above prescription also yields directly the insertion of the momentum necessary if the +interaction contains derivatives. +In Section 14.2, the Feynman rule for a g-loop n-point fundamental vertex of states +(V1, . . . , Vn) was found to be given by (14.33): +Vg,n(V1, . . . , Vn) = +� +Rg,n +ωg,n +Mg,n(V1, . . . , Vn) = +(15.21) +where Rg,n is a section over the fundamental region Vg,n ⊂ Mg,n (12.42b) which cannot be +covered from the plumbing fixture of lower-dimensional surfaces. +From the example Example 15.1, it should be clear that the g-loop n-point contribution +to the action can be obtained simply by replacing every state with a string field in Vg,n: +Sg,n = ℏg g2g−2+n +s +n! +Vg,n(Ψn). +(15.22) +where Ψn := Ψ⊗n. +The power of the coupling constant has been reinstated: it can be +motivated by the fact that it should have the same power as the corresponding amplitude +(Section 3.1.1). Note that the interactions are defined only when the power of gs is positive: +229 + +χg,n = 2 − 2g − n < 0. We have also written explicitly the power of ℏ, which counts the +number of loops. +Before closing this section, we need to comment on the effect of the constraints (15.14) +on the interactions. Building a Feynman graph by gluing two m- and n-point interactions +with a propagator, one finds that the states proportional to φx,r for x ̸=↓↓ do not propagate +inside internal legs +Vg,m(V1, . . . , Vm−1, φr)⟨φc +r| b+ +0 b− +0 +1 +L+ +0 +|φc +s⟩ Vg′,n(W1, . . . , Wn−1, φs) += Vg,m(V1, . . . , Vm−1, φ↓↓,r)⟨φc +↓↓,r| b+ +0 b− +0 +1 +L+ +0 +|φc +↓↓,s⟩ Vg′,n(W1, . . . , Wn−1, φ↓↓,s). +(15.23) +Thus, they do not contribute to the final result even if the interactions contain them. While +the conditions L− +0 = b− +0 = 0 were found to be necessary for defining off-shell amplitudes, the +condition b+ +0 = 0 does not arise from any consistency requirement. But, it is also consistent +with the interactions, since only fundamental vertices have a chance to give a non-vanishing +result for states which do not satisfy (15.14). Hence, the interactions (15.22) are compatible +with the definition of the kinetic term and the restriction of the string field. +15.2.3 +Action +The interacting gauge-fixed action is built from the kinetic term V0,2 (15.17) and from the +interactions Vg,n (15.22) with χg,n < 0. However, this is not sufficient: we have seen in +Section 14.3 that it makes sense to consider the vertices with χg,n ≥ 0. First, we should +consider the 1-loop cosmological constant V1,0. Then, we can also add the classical source +V0,1 and the tree-level cosmological constant V0,0. With all the terms together, the action +reads: +S = +� +g,n≥0 +ℏg g2g−2+n +s +n! +Vg,n(Ψn) +:= 1 +2 ⟨Ψ| c− +0 c+ +0 L+ +0 δL− +0 ,0 |Ψ⟩ + +�′ +g,n≥0 +ℏg g2g−2+n +s +n! +Vg,n(Ψn). +(15.24) +where Vn was defined in (14.58). A prime on the sum indicates that the term g = 0, n = 2 +is removed, such that one can single out the kinetic term. We will often drop the delta +function imposing L− +0 = 0 because the field are taken to satisfy this constraint. +Rewriting the vertices in terms of the products ℓg,n defined in (14.65) +Vg,n(Ψn) :=⟨Ψ| c− +0 +��ℓg,n−1(Ψn−1) +� +(15.25) +leads to the alternative form +S = +� +g,n≥0 +ℏg g2g−2+n +s +n! +⟨Ψ| c− +0 +��ℓg,n−1(Ψn−1) +� +. +(15.26) +The definition (14.56) leads to the following explicit expression for ℓ0,1: +ℓ0,1(Ψcl) = c+ +0 L+ +0 |Ψcl⟩ . +(15.27) +In most cases, the terms g = 0, n = 0, 1 vanish such that the action reads: +S = +� +g,n≥0 +χg,n≤0 +ℏg g2g−2+n +s +n! +Vg,n(Ψn). +(15.28) +230 + +However, we will often omit the condition χg,n ≤ 0 to simplify the notation, except when the +distinction is important, and the reader can safely assumes V0,0 = V0,1 = 0 if not otherwise +stated. The classical action is obtained by setting ℏ = 0: +Scl = 1 +2 ⟨Ψcl| c− +0 c+ +0 L+ +0 |Ψcl⟩ + +� +n≥3 +gn +s +n! V0,n(Ψn +cl). +(15.29) +Rescaling the string field by g−1 +s +gives the more canonical form of the action (using the +same symbol): +S = +� +g,n≥0 +ℏgg2g−2 +s +1 +n! Vg,n(Ψn) +:= +1 +2g2s +⟨Ψ| c− +0 c+ +0 L+ +0 δL− +0 ,0 |Ψ⟩ + 1 +g2s +�′ +g,n≥0 +(ℏg2 +s)g +n! +Vg,n(Ψn). +(15.30) +In the path integral, the action is divided by ℏ such that +S +ℏ = +� +g,n≥0 +(ℏg2 +s)g−1 1 +n! Vg,n(Ψn). +(15.31) +This shows that there is a single coupling constant ℏg2 +s, instead of two (ℏ and gs separately) +as it looks at the first sight. This makes sense because gs is in fact the expectation value +of the dilaton field (2.166) and its value can be changed by deforming the background with +dilatons [16, 17, 197]. +The previous remark also allows to easily change the normalization of the action, for +example, to perform a Wick rotation, to normalize canonically the action in terms of space- +time fields, or reintroduce ℏ. Rescaling the action by α is equivalent to rescale g2 +s by α−1: +S → α S +=⇒ +g2 +s → g2 +s +α . +(15.32) +The linearized equation of motion is: +L+ +0 |Ψ⟩ = 0, +(15.33) +which corresponds to the Siegel gauge equation of motion of the free theory (10.116). Hence, +this equation is not sufficient to determine the physical states (cohomology of the BRST op- +erator, Chapter 8), as discussed in Chapter 10, and additional constraints must be imposed. +One can interpret this by saying that the action (15.24) provides only the Feynman rules, +not the physical states. Removing the gauge fixing will be done in Sections 15.3 and 15.4. +The action (15.24) looks overly more complicated than a typical QFT theory: instead +of few interaction terms for low n (n ≤ 4 in d = 4 renormalizable theories), it has contact +interactions of all orders n ∈ N. The terms with g ≥ 1 are associated to quantum corrections +as indicate the power of ℏ, which means that they can be interpreted as counter-terms. But, +how is it that one needs counter-terms despite the claim that every Feynman graphs (includ- +ing the fundamental vertices) in SFT are finite? The role of renormalization is not only to +cure UV divergences, but also IR divergences (due to vacuum shift and mass renormaliza- +tion). Equivalently, this can be understood by the necessity to correct the asymptotic states +of the theory, or to consider renormalized instead of bare quantities. Indeed, the asymptotic +states obtained from the linearized classical equations of motion are idealization: turning +on interactions modify the states. In typical QFTs, these corrections are infinite and renor- +malization is crucial to extract a number; however, even if the effect is finite, it is needed to +describe correctly the physical quantities [250, p. 411]. There is a second reason for these +231 + +additional terms: when relaxing the gauge fixing condition, the path integral is anomalous +under the gauge symmetry, and the terms with g > 0 are necessary to cancel the anomaly +(this will be discussed more precisely in Section 15.4). It may thus seem that SFT cannot +be predictive because of the infinite number of counter-terms. Fortunately, this is not the +case: the main reason for the loss of predictability in non-renormalizable theory is that +the renormalization procedure introduces an infinite number2 of arbitrary parameters (and +thus making a prediction would require to have already made an infinite number of observa- +tions to determine all the parameters). These parameters come from the subtraction of two +infinities: there is no unique way to perform it and thus one needs to introduce a new para- +meter. The case of SFT is different: since every quantity is finite, the renormalization has +no ambiguity because one subtracts two finite numbers, and the result is unambiguous. As +a consequence, renormalization does not introduce any new parameter and there is a unique +coupling constant gs in the theory, which is determined by the tree-level cubic interaction. +The coupling constants of higher-order and higher-loop interactions are all determined by +powers of gs, and thus a unique measurement is sufficient to make predictions. +Another important point is that the action (15.24) is not uniquely defined. The definition +of the vertices depends on the choice of the local coordinates and of the stub parameter s0. +Changing them modifies the vertices, and thus the action. +But, one can show that the +different theories are related by field redefinitions and are thus equivalent. +15.3 +Classical gauge invariant theory +In the previous section, we have found the gauge fixed action (15.24). Since the complete +gauge invariant quantum action has a complicated structure, it is instructive to first focus +on the classical action (15.29). The full action is discussed in Section 15.4. +The gauge fixing is removed by relaxing the b+ +0 = 0 constraint on the field (the other +constraints must be kept in order to have well-defined the interactions). The classical field +Ψcl is then defined by: +Ψcl ∈ H− ∩ ker L− +0 , +Ngh(Ψcl) = 2. +(15.34) +The restriction on the ghost number translates the condition that the field is classical, i.e. +that there are no spacetime ghosts at the classical level. The relation (15.6) implies that all +components have vanishing spacetime ghost number. +In the free limit, the gauge invariant action should match (10.105) +S0,2 = 1 +2 ⟨Ψ| c− +0 QB |Ψ⟩ . +(15.35) +and lead to the results from Section 10.5. A natural guess is that the form of the interactions +is not affected by the gauge fixing (the latter usually modifies the propagator but not the +interactions). This leads to the gauge invariant classical action: +Scl = 1 +2 ⟨Ψcl| c− +0 QB |Ψcl⟩ + 1 +g2s +� +n≥3 +gn +s +n! V0,n(Ψn +cl), +(15.36) +where the vertices V0,n with n ≥ 3 are the ones defined in (14.33) (we consider the case +where V0,0 = V0,1 = 0). It is natural to generalize the definition of V0,2 as: +V0,2(Ψ2 +cl) :=⟨Ψcl| c− +0 QB |Ψcl⟩ +(15.37) +2In practice, this number does not need to be infinite to wreck predictability, it is sufficient that it is +very large. +232 + +such that +Scl = 1 +g2s +� +n≥2 +gn +s +n! V0,n(Ψn +cl) = 1 +g2s +� +n≥2 +gn +s +n! ⟨Ψcl| c− +0 +��ℓ0,n−1(Ψn−1 +cl +) +� +, +(15.38) +where (15.37) implies: +ℓ0,1(Ψcl) = QB |Ψcl⟩ . +(15.39) +The equation of motion is +Fcl(Ψcl) := +� +n≥1 +gn−1 +s +n! +ℓ0,n(Ψn +cl) = QB |Ψcl⟩ + +� +n≥2 +gn−1 +s +n! +ℓ0,n(Ψn +cl) = 0. +(15.40) +Computation – Equation (15.40) +δScl = 1 +g2s +� +n≥2 +gn +s +n! n{δΨcl, Ψn−1 +cl +}0 = 1 +g2s +� +n≥2 +gn +s +(n − 1)! ⟨δΨcl| c− +0 +��ℓ0,n−1(Ψn−1 +cl +� +. +(15.41) +The first equality follows because the vertex is completely symmetric. Simplifying and +shifting n, one obtains c− +0 |Fcl⟩. The factor c− +0 is invertible because of the constraint +b− +0 = 0 imposed on the field. +The action is invariant +δΛScl = 0 +(15.42) +under the gauge transformation +δΛΨcl = +� +n≥0 +gn +s +n! ℓ0,n+1(Ψn +cl, Λ) = QB |Λ⟩ + +� +n≥1 +gn +s +n! ℓ0,n+1(Ψn +cl, Λ). +(15.43) +The gauge algebra is [262, sec. 4]: +[δΛ2, δΛ1]Ψcl = δΛ(Λ1,Λ2,Ψcl) |Ψcl⟩ + +� +n≥0 +gn+2 +s +n! +ℓ0,n+3 +� +Ψn +cl, Λ2, Λ1, Fcl(Ψcl) +� +, +(15.44a) +where Fcl is the equation of motion (15.40), and Λ(Λ1, Λ2, Ψcl) is a field-dependent gauge +parameter: +Λ(Λ1, Λ2, Ψcl) = +� +n≥0 +gn+1 +s +n! +ℓ0,n+2(Λ1, Λ2, Ψn +cl) += gs ℓ0,2(Λ1, Λ2) + +� +n≥1 +gn+1 +s +n! +ℓ0,n+2(Λ1, Λ2, Ψn +cl). +(15.44b) +The classical gauge algebra is complicated which explains why a direct quantization (for +example through the Faddeev–Popov procedure) cannot work: the second term in (15.44a) +indicates that the algebra is open (it closes only on-shell), while the first term is a gauge +transformation with a field-dependent parameter. As reviewed in Appendix C.3, both prop- +erties require using the BV formalism for the quantization, and the latter is performed in +Section 15.4. An important point is that if the theory had only cubic interactions, i.e. if +∀n ≥ 4 : +V0,4(V1, . . . , Vn) = 0, +ℓg,n−1(V1, . . . , Vn−1) = 0, +(cubic theory), (15.45) +then the algebra closes off-shell and Λ(Λ1, Λ2, Ψcl) becomes field-independent. +233 + +Computation – Equation (15.42) +δΛScl = +� +n≥2 +gn−2 +s +n! +nV0,n(δΨcl, Ψn−1 +cl +) = +� +m,n≥0 +gm+n−1 +s +m! n! +V0,n+1 +� +ℓ0,m+1(Ψm +cl , Λ), Ψn +cl +� += +� +m≥0 +m +� +n=0 +gm−1 +s +(m − n)! n! +� +ℓm−n+1(Ψm−n +cl +, Λ) +�� c− +0 |ℓ0,n(Ψn +cl)⟩ . +For simplicity we have extended the sum up to n = 0 and m = 0 by using the fact that +lower-order vertices vanish. The bracket can be rewritten as += ⟨ℓ0,n(Ψn +cl)| c− +0 +��ℓ0,m−n+1(Ψm−n +cl +, Λ) +� += V0,m−n+2 +� +ℓ0,n(Ψn +cl), Ψm−n +cl +, Λ +� += −V0,m−n+2 +� +Λ, ℓ0,n(Ψn +cl), Ψm−n +cl +� += ⟨Λ| c− +0 +��ℓ0,m−n+1 +� +ℓ0,n(Ψn +cl), Ψm−n +cl +�� +. +Then, one needs to use the identity (defined for all m ≥ 0) +0 = +m +� +n=0 +m! +(m − n)! n! ℓ0,m−n+1 +� +ℓ0,n(Ψn +cl), Ψm−n +cl +� +, +(15.46) +which comes from (14.70). Multiplying this by gm−1 +s +/m! and summing over m ≥ 0 +proves (15.42). +Remark 15.1 (L∞ algebra) The identities satisfied by the products ℓ0,n from the gauge +invariance of the action implies that they form a L∞ homotopy algebra [74, 169, 262] (for +more general references, see [106, 108, 148, 149]). The latter can also be mapped to a BV +structure, which explains why the BV quantization Section 15.4 is straightforward. This +interplay between gauge invariance, covering of the moduli space, BV and homotopy algebra +is particularly beautiful. It has also been fruitful in constructing super-SFT. +15.4 +BV theory +As indicated in the previous section (Section 15.3), the classical gauge algebra is open and +has field-dependent structure constants. The BV formalism (Appendix C.3) is necessary to +define the theory. +In the BV formalism, the classical action for the physical fields is extended to the +quantum master action by solving the quantum master equation (C.114). +It is generic- +ally difficult to build this action exactly, but the discussion of Section 10.3 can serve as a +guide: it was found that the free quantum action (with the tower of ghosts) has exactly the +same form as the free classical action (without ghosts). Hence, this motivates the ansatz +that it should be of the same form as the classical action (15.36) to which are added the +234 + +counter-terms from (15.24): +S = 1 +g2s +� +g≥0 +ℏgg2g +s +� +n≥0 +gn +s +n! Vg,n(Ψn) +(15.47a) += 1 +2 ⟨Ψ| c− +0 QB |Ψ⟩ + +�′ +g,n≥0 +ℏgg2g−2+n +s +n! +Vg,n(Ψn) +(15.47b) += 1 +g2s +� +g,n≥0 +ℏgg2g−2+n +s +n! +⟨Ψ| c− +0 +��ℓg,n−1(Ψn−1) +� +, +(15.47c) +but without any constraint on the ghost number of Ψ: +Ψ ∈ H− ∩ ker L− +0 . +(15.48) +In order to show that (15.47) is a consistent quantum master action, it is necessary to +show that it solves the master BV equation (C.114): +(S, S) − 2ℏ∆S = 0. +(15.49) +The first step is to introduce the fields and antifields. In fact, because the CFT ghost number +induces a spacetime ghost number, there is a natural candidate set. +The string field is expanded as (15.1) +|Ψ⟩ = +� +r +ψr |φr⟩ , +(15.50) +where the {φr} forms a basis of H−. The string field can be further separated as: +Ψ = Ψ+ + Ψ−, +(15.51) +where Ψ− (Ψ+) contains only states which have negative (positive) cylinder ghost numbers +(this gives an offset of 3 when using the plane ghost number): +Ψ− = +� +r +� +nr≤2 +|φr⟩ ψr, +Ψ+ = +� +r +� +ncr>2 +b− +0 |φc +r⟩ ψ∗ +r. +(15.52) +The order of the basis states and coefficients matter if they anti-commute. The sum in Ψ+ +can be rewritten as a sum over nr ≤ 2 like the first term since nr +nc +r = 6. Correspondingly, +the spacetime ghost numbers (15.6) for the coefficients in Ψ− (Ψ+) are positive (negative) +G(ψr) ≥ 0, +G(ψ∗ +r) < 0. +(15.53) +Moreover, one finds that the ghost numbers of ψr and ψ∗ +r are related as: +G(ψ∗ +r) = −1 − G(ψr), +(15.54) +which also implies that they have opposite parity. Comparing with Appendix C.3, this shows +that the ψr (ψ∗ +r) contained in Ψ− (Ψ+) can be identified with the fields (antifields). +Computation – Equation (15.54) +G(ψ∗ +r) = 2 − Ngh(b− +0 φc +r) = 2 + 1 − nc +r += 3 − (6 − nr) = −3 + (2 − G(ψr)) = −1 − G(ψr). +235 + +In terms of fields and antifields, the master action is +∂RS +∂ψr +∂LS +∂ψ∗r ++ ℏ ∂R∂LS +∂ψr∂ψ∗r += 0. +(15.55) +Plugging the expression (15.47) of S inside and requiring that the expression vanishes order +by order in g and n give the set of equations: +� +g1,g2≥0 +g1+g2=g +� +n1,n2≥0 +n1+n2=n +∂RSg1,n1 +∂ψr +∂LSg2,n2 +∂ψ∗r ++ ℏ ∂R∂LSg−1,n +∂ψr∂ψ∗r += 0, +(15.56) +where Sg,n was defined in (15.22). This holds true due to the identity (14.70) (the complete +proof can be found in [262, pp. 42–45]). The fact that the second term is not identically zero +means that the measure is not invariant under the classical gauge symmetry (anomalous +symmetry): corrections need to be introduced to cancel the anomaly. It is a remarkable +fact that one can construct directly the quantum master action in SFT and that it takes +the same form as the classical action. +15.5 +1PI theory +The BV action is complicated: instead, it is often simpler and sufficient to work with the +1PI effective action. The latter incorporates all the quantum corrections in 1PI vertices +such that scattering amplitudes are expressed only in terms of tree Feynman graphs (there +are no loops in diagrams since they correspond to quantum effects, already included in the +definitions of the vertices). +A 1PI graph is a Feynman graph which stays connected if one cuts any single internal +line. On the other hand, a 1PR graph splits in two disconnected by cutting one of the line. +The scattering amplitudes Ag,n are built by summing all the different ways to connect two +1PI vertices with a propagator: diagrams connecting two legs of the same 1PI vertex are +forbidden by definition. +The g-loop n-point 1PR and 1PI Feynman diagrams are associated to some regions of +the moduli space Mg,n. Comparing the previous definitions with the gluing of Riemann +surfaces (Section 12.3), 1PR diagrams are obtained by gluing surfaces with the separating +plumbing fixture (Section 14.1.1). Thus, the 1PR and 1PI regions F1PR +g,n +and V1PI +g,n can be +identified with the regions defined in (12.43a) and (12.43b). In particular, the n-point 1PI +interaction is the sum over g of the g-loop n-point 1PI interactions (14.63): +V1PI +n +(V1, . . . , Vn) := +:= +� +g≥0 +(ℏg2 +s)g V1PI +g,n (V1, . . . , Vn), +V1PI +g,n (V1, . . . , Vn) := +� +R1PI +g,n +ωg,n +Mg,n(V1, . . . , Vn), +(15.57) +where R1PI +g,n is a section of Pg,n over V1PI +g,n . +Given the interactions vertices, it is possible to follow the same reasoning as in Sec- +tions 15.2 and 15.3. +236 + +The gauge fixed 1PI effective action reads: +S1PI = 1 +g2s +� +n≥0 +gn +s +n! V1PI +n +(Ψn) := 1 +2 ⟨Ψ| c− +0 c+ +0 L+ +0 |Ψ⟩ + 1 +g2s +�′ +n≥0 +gn +s +n! V1PI +n +(Ψn). +(15.58) +Here, the prime means again that the term g = 0, n = 2 is excluded from the definition of +V1PI +2 +. The action has the same form as the classical gauge fixed action (15.29), which is +logical since it generates only tree-level Feynman graphs. For this reason the vertices V1PI +n +have exactly the same properties as the brackets V0,n. This fact can be used to write the +1PI gauge invariant action: +S1PI = 1 +2 ⟨Ψ| c− +0 QB |Ψ⟩ + 1 +g2s +�′ +n≥0 +gn +s +n! V1PI +n +(Ψn), +(15.59) +which mirrors the classical gauge invariant action (15.36). Then it is straightforward to see +that it enjoys the same gauge symmetry upon replacing the tree-level vertices by the 1PI +vertices. But, since this action incorporates all quantum corrections this also proves that +the quantum theory is correctly invariant under a quantum gauge symmetry. +Remark 15.2 The 1PI action (15.59) can also be directly constructed from the BV action +(15.47). +15.6 +Suggested readings +• Gauge fixed and classical gauge invariant closed SFT [262] (see also [138, 139]). +• BV closed SFT [262] (see also [245]). +• Construction of the open–closed BV SFT [264]. +• 1PI SFT [213, 214, 42, sec. 4.1, 5.2]. +237 + +Chapter 16 +Background independence +Abstract +Spacetime background independence is a fundamental property of any candidate +quantum gravity theory. In this chapter, we outline the proof of background independence +for the closed SFT by proving that the equations of motion of two background related by a +marginal deformation are equivalent after a field redefinition. +16.1 +The concept of background independence +Background independence means that the formalism does not depend on the background +– if any – used to write the theory. A dependence in the background would imply that +there is a distinguished background among all possibilities, which seems in tension with the +dynamics of spacetime and the superposition principle from quantum mechanics. Moreover, +one would expect a fundamental theory to tell which backgrounds are consistent and that +they could be derived instead of postulated. Background independence allows spacetime to +emerge as a consequence of the dynamics of the theory and of its defining fundamental laws. +Background independence can be manifest or not. In the second case, one needs to fix a +background to define the theory, but the dynamics on different backgrounds are physically +equivalent.1 This implies that two theories with different backgrounds can be related, for +example by a field redefinition. +While fields other than the metric can also be expanded around a background, no diffi- +culty is expected in this case. Indeed, the topic of background independence is particularly +sensible only for the metric because it provides the frame for all other computations – and +in particular for the questions of dynamics and quantization. Generally, these questions are +subsumed into the problem of the emergence of time in a generally covariant theory. In +the previous language, QFTs without gravity are (generically) manifestly background inde- +pendent after minimal coupling.2 For example, a classical field theory is defined on a fixed +Minkowski background and a well-defined time is necessary to perform its quantization and +to obtain a QFT, but it is not needed to choose a background for the other fields. For this +reason, the extension of a QFT on a curved background is generally possible if the space- +time is hyperbolic, implying that there is a distinguished time direction. But the coupling +to gravity is difficult and restricted to a (semi-)classical description. +What is the status of background independence in string theory? The worldsheet for- +mulation requires to fix a background (usually Minkowski) to quantize the theory and to +compute scattering amplitudes. Thus, the quantum theory is at least not manifestly back- +1This does not mean that the physics in all backgrounds are identical, but that the laws are. Hence, a +computation made in one specific background can be translated into another background. +2However, non-minimal coupling terms may be necessary to make the theory physical. +238 + +ground independent. On the other hand, the worldsheet action can be modified to a generic +CFT including a generic non-linear sigma model describing an arbitrary target spacetime. +Conformal invariance reproduces (at leading order) Einstein equations coupled to various +matter and gauge field equations of motion. From this point of view, the classical theory can +be written as a manifestly background independent theory, and this provides hopes that the +quantum theory may be also background independent, even if non-manifestly. This idea is +supported by other definitions of string theory (e.g. through the AdS/CFT conjecture – and +other holographic realizations – or through matrix models) which provide, at least partially, +background independent formulations. +Ultimately, the greatest avenue to establish the background independence is string field +theory. Indeed, the form of the SFT action and of its properties (gauge invariance, equa- +tion of motion. . . ) are identical irrespective of the background [234]. This provides a good +starting point. The background dependence enters in the precise definition of the string +products (BRST operator and vertices). The origin of this dependence lies in the derivation +of the action (Chapters 14 and 15): one begins with a particular CFT describing a given +background (spacetime compactifications, fluxes, etc.) and defines the vertices from correl- +ation functions of vertex operators, and the Hilbert space from the CFT operators. As a +consequence, even though it is clear that no specific property of the background has been +used in the derivation – and that the final action describes SFT for any background –, this +is not sufficient to establish background independence. Since the theory assumes implicitly +a background choice, one cannot guarantee that the physical quantities have no residual +dependence in the background, even if the action looks superficially background independ- +ent. Background independence in SFT is thus the statement that theories characterized by +different CFTs can be related by a field redefinition. +In this chapter, we will sketch the proof of background independence for backgrounds +related by marginal deformations.3 It is possible to prove it at the level of the action [232, +233], or at the level of the equations of motion [226]. The advantage of the second approach +is that one can use the 1PI theory, which simplifies vastly the analysis. It also generalizes +directly to the super-SFT. +Remark 16.1 (Field theory on the CFT space) As mentioned earlier, the string field +is defined as a functional on the state space of a given CFT and not as a functional on the +field theory space (off-shell states would correspond to general QFTs, only on-shell states are +CFTs). In this case, background independence would amount to reparametrization invariance +of the action in the theory space, and would thus almost automatically hold. A complete +formulation of SFT following this line is currently out of reach, but some ideas can be found +in [255]. +16.2 +Problem setup +Given a SFT on a background, there are two ways to describe it on another background: +• deform the worldsheet CFT and express the SFT on the new background; +• expand the original action around the infinitesimal classical solution (to the linearised +equations of motion) corresponding to the deformation. +Background independence amounts to the equivalence of both theories up to a field redefin- +ition. The derivation can be performed at the level of the action or of the equations of +motion. To prove the background independence at the quantum level, one needs to take +into account the changes in the path integral measure or to work with the 1PI action. +3An alternative approach based on morphism of L∞ algebra is followed in [169]. +239 + +The simplest case is when the two CFTs are related by an infinitesimal marginal deform- +ation +δScft = λ +2π +� +d2z ϕ(z, ¯z), +(16.1) +with ϕ a (1, 1)-primary operator and λ infinitesimal. The two CFTs are denoted by CFT1 +and CFT2, and quantities associated to each CFTs is indexed with the appropriate number. +Establishing background independence in this case also implies it for finite marginal +deformation since they can be built from a series of successive deformations. In the latter +case, the field redefinition may be singular, which reflects that the parametrization of one +CFT is not adapted for the other (equivalently, the coordinate systems for the string field +breaks down), which is expected if both CFTs are far in the field theory space. +Remember the form of the 1PI action (15.59): +S1[Ψ1] = 1 +g2s +� +�1 +2 ⟨Ψ1| c− +0 QB |Ψ1⟩ + +�′ +n≥0 +1 +n! V1PI +n +(Ψn +1) +� +� , +(16.2) +where the prime indicates that vertices with n < 3 do not include contributions from the +sphere. In all this chapter, we remove the index 1PI to lighten the notations. The equation +of motion is: +F1(Ψ1) = QB |Ψ1⟩ + +� +n +1 +n! ℓn(Ψn +1) = 0. +(16.3) +16.3 +Deformation of the CFT +Consider the case where the theory CFT1 is described by an action Scft,1[ψ1] given in terms +of fields ψ1. Then, the deformation of this action by (16.1) gives an action for CFT2: +Scft,2[ψ1] = Scft,1[ψ1] + λ +2π +� +d2z ϕ(z, ¯z). +(16.4) +Correlation functions on a Riemann surface Σ in both theories can be related by expanding +the action to first order in λ in the path integral: +�� +i +Oi(zi, ¯zi) +� +2 += +� +exp +� +− λ +2π +� +d2z ϕ(z, ¯z) +� � +i +Oi(zi, ¯zi) +� +1 +(16.5a) +≈ +�� +i +Oi(zi, ¯zi) +� +1 +− λ +2π +� +Σ +d2z +� +ϕ(z, ¯z) +� +i +Oi(zi, ¯zi) +� +1 +, +(16.5b) +where the Oi are operators built from the matter fields ψ1. This expression presents two +obvious problems. First, the correlation function may diverge when ϕ collides with one of the +insertions, i.e. when z = zi in the integration. Second, there is an inherent ambiguity: the +correlation functions are written in terms of operators in the Hilbert space of CFT1, which +is different from the CFT2 Hilbert space, and there is no canonical isomorphism between +both spaces. +Seeing the Hilbert space as a vector bundle over the CFT theory space, the second +problem can be solved by introducing a connection on this bundle. This allows to relate +Hilbert spaces of neighbouring CFTs. In fact, the choice of a non-singular connection also +regularizes the divergences. +The simplest definition of a connection corresponds to cut unit disks around each operator +insertions [30, 198, 199, 210, 238]. This amounts to define the variation between the two +240 + +correlation functions as: +δ +�� +i +Oi(zi, ¯zi) +� +1 += − λ +2π +� +Σ−∪iDi +d2z +� +ϕ(z, ¯z) +� +i +Oi(zi, ¯zi) +� +1 +. +(16.6) +The integration is over Σ minus the disks Di = {|wi| ≤ 1} where wi is the local coordinate +for the insertion Oi. The divergences are cured because ϕ never approaches another operator +since the corresponding regions have been removed. The changes in the correlation functions +induce a change in the string vertices denoted by δVn(V1, . . . , Vn). +The next step consists in computing the deformations of the operator modes. Since it +involves only a matter operator, the modes in the ghost sector are left unchanged. The +Virasoro generators change as: +δLn = λ +� +|z|=1 +d¯z +2πi zn+1ϕ(z, ¯z), +δ ¯Ln = λ +� +|z|=1 +dz +2πi ¯zn+1ϕ(z, ¯z). +(16.7) +As a consequence, the BRST operator changes as +δQB = λ +� +|z|=1 +d¯z +2πi c(z)ϕ(z, ¯z) + λ +� +|z|=1 +d¯z +2πi ¯c(¯z)ϕ(z, ¯z). +(16.8) +One can prove that +{QB, δQB} = O(λ2) +(16.9) +such that the BRST charge QB + δQB in CFT2 is correctly nilpotent if QB is nilpotent in +CFT1. +For the deformation to provide a consistent SFT, the conditions b− +0 = 0 and L− +0 = 0 must +be preserved. The first is automatically satisfied since the ghost modes are not modified. +Considering an weight-(h, h) operator O, one finds +δL− +0 |O⟩ = λ +� +|z|=1 +d¯z +2πi z +� +p,q +zp−1¯zq−1 |Op,q⟩ − λ +� +|z|=1 +d¯z +2πi +� +p,q +zp−1¯zq−1 |Op,q⟩ , +(16.10) +where Op,q are the fields appearing in the OPE with ϕ: +ϕ(z, ¯z)O(0, 0) = +� +p,q +zp−1¯zq−1Op,q(0, 0). +(16.11) +The terms with p ̸= q vanish because the contour integrals are performed around circles of +unit radius centred at the origin. Moreover, the terms p = q are identical and cancel with +each other, showing that δL− +0 = 0 when acting on states satisfying L− +0 = 0. +The SFT action S2[Ψ1] in the new background reads +S2[Ψ1] = S1[Ψ1] + δS1[Ψ1] +(16.12) +where the change δS1 in the action is induced by the changes in the string vertices: +δS1[Ψ1] = 1 +g2s +� +�1 +2 ⟨Ψ1| c− +0 δQB |Ψ1⟩ + +� +n≥0 +1 +n! δVn(Ψn +1) +� +� . +(16.13) +The equation of motion is: +F2(Ψ1) = F1(Ψ1) + λ δF1(Ψ1) = 0, +(16.14) +where F1 is given in (16.3) and +λ δF1(Ψ1) = δQB |Ψ1⟩ + +� +n +1 +n! δℓn(Ψn +1). +(16.15) +241 + +16.4 +Expansion of the action +Given a (1, 1) primary ϕ, a BRST invariant operator is c¯cϕ. Hence the field +|Ψ1⟩ = λ |Ψ0⟩ , +|Ψ0⟩ = c1¯c1(0) |ϕ⟩ +(16.16) +is a classical solution to first order in λ since the interactions on the sphere are at least +cubic. +Separating the string field as the contribution from the (fixed) background and a fluctu- +ation Ψ′ +|Ψ1⟩ = λ |Ψ0⟩ + |Ψ′⟩ , +(16.17) +the action expanded to first order in λ reads: +S1[Ψ1] = S1[Ψ0] + S′[Ψ′], +(16.18) +where +S′[Ψ′] = 1 +g2s +� +1 +2 ⟨Ψ′| c− +0 QB |Ψ′⟩ + +� +n +1 +n! +� +Vn(Ψ′n) + λ Vn+1(Ψ0, Ψ′n) +� +� +. +(16.19) +The equation of motion is: +F′(Ψ′) := F1(Ψ′) + λ δF′(Ψ′) = 0, +(16.20) +where F1 is given in (16.3) and +δF′(Ψ′) = +� +n +1 +n! ℓn+1(Ψ0, Ψ′n). +(16.21) +16.5 +Relating the equations of motion +In the previous section, we have derived the equations of motion for two different descriptions +of a SFT obtained after shifting the background: (16.14) arises by deforming the CFT and +computing the changes in the BRST operator and string products, while (16.20) arises by +expanding the SFT action around the new background. The theory is background independ- +ent if both sets of equations (16.14) and (16.20) are related by a (possibly field-dependent) +linear transformation M(Ψ′) after a field redefinition of Ψ1 = Ψ1(Ψ′): +F1(Ψ1) + λ δF1(Ψ1) = +� +1 + λM(Ψ′) +�� +F1(Ψ′) + λ δF′(Ψ′) +� +, +(16.22a) +|Ψ1⟩ = |Ψ′⟩ + λ |δΨ′⟩ . +(16.22b) +The zero-order equation is automatically satisfied. To first order, this becomes +d +dλF1(Ψ′ + λδΨ′) +���� +λ=0 ++ δF1(Ψ1) − δF′(Ψ′) = M(Ψ′)F1(Ψ′). +(16.23) +Taking Ψ′ to be a solution of the original action removes the RHS, such that: +λ QB |δΨ′⟩ + λ +� +n +1 +n! ℓn+1(δΨ′, Ψ′n) + δQB |Ψ′⟩ ++ +� +n +1 +n! δℓn(Ψ′n) − λ +� +n +1 +n! ℓn+1(Ψ0, Ψ′n) = 0. +(16.24) +242 + +To simplify the computations, it is simpler to consider the inner product of this quantity +with an arbitrary state A (assumed to be even): +∆ := λ ⟨A| c− +0 QB |δΨ′⟩ + λ +� +n +1 +n! Vn+2(A, δΨ′, Ψ′n) +⟨A| c− +0 δQB |Ψ′⟩ ++ +� +n +1 +n! δVn+1(A, Ψ′n) − λ +� +n +1 +n! Vn+2(A, Ψ0, Ψ′n). +(16.25) +The goal is to prove the existence of δΨ′ such that ∆ = 0 up to the zero-order equation of +motion F1(Ψ′) = 0. +16.6 +Idea of the proof +In this section, we give an idea of how the proof ends, referring to [226] for the details. +The first step is to introduce new vertices V′ +0,3 and V′ +n parametrizing the variations of +the string vertices: +⟨A| c− +0 δQB |B⟩ = λ V′ +0,3(Ψ0, B, A), +δVn(Ψ′n) = λ V′ +n+1(Ψ0, Ψ′n), +(16.26) +where the notation (13.19) has been used. Each subspace V′ +g,n is defined such that the LHS +is recovered upon integrating the appropriate ωg,n over this section segment. Next, the field +redefinition δΨ′ is parametrized as: +⟨A| c− +0 |δΨ′⟩ = +� +n +1 +n! Bn+2(Ψ0, Ψ′n, A). +(16.27) +The objective is to prove the existence (and if possible the form) of the subspaces Bn+2. +Both the vertices V′ +n and Bn admit a genus expansion: +V′ +n = +� +g≥0 +V′ +g,n, +Bn = +� +g≥0 +Bg,n. +(16.28) +Plugging the new expressions in (16.25) give: +∆ = − +� +n +1 +n! Bn+2(Ψ0, Ψ′n, QBA) + +� +m,n +1 +m!n! Bn+2(Ψ0, Ψ′m, ℓn+1(A, Ψ′n)) ++ +� +n +1 +n! V′ +n+2(A, Ψ0, Ψ′n) − +� +n +1 +n! Vn+2(A, Ψ0, Ψ′n). +(16.29) +Next, the BRST identity (13.46) and the equation of motion F1(Ψ′) = 0 allow to rewrite +the first term as: +Bn+2(Ψ0, Ψ′n, QBA) = ∂Bn+2(Ψ0, Ψ′n, A) + n Bn+2(Ψ0, Ψ′n−1, QBΨ′, A) +(16.30a) += ∂Bn+2(Ψ0, Ψ′n, A) − +� +m +n +m! Bn+2(Ψ0, Ψ′n−1, ℓm(Ψ′m), A). +(16.30b) +In the second term, the sum over n is shifted. Combining everything together gives: +∆ = +� +n +1 +n! ∂Bn+2(Ψ0, Ψ′n, A) − +� +m,n +1 +m!n! Bn+3(Ψ0, Ψ′n, ℓm(Ψ′m), A) ++ +� +m,n +1 +m!n! Bn+2(Ψ0, Ψ′m, ℓn+1(A, Ψ′n)) + +� +n +1 +n! V′ +n+2(A, Ψ0, Ψ′n) +− +� +n +1 +n! Vn+2(A, Ψ0, Ψ′n). +(16.31) +243 + +Solving for ∆ = 0 requires that each term with a different power of Ψ′ vanishes independ- +ently: +∂Bn+2(Ψ0, Ψ′n, A) = − V′ +n+2(A, Ψ0, Ψ′n) + Vn+2(A, Ψ0, Ψ′n) ++ +� +m1,m2 +m1+m2=n +n! +m1!m2! Bm1+3(Ψ0, Ψ′m1, ℓm2(Ψ′m2), A) +− +� +m1,m2 +m1+m2=n +n! +m1!m2! Bm1+2 +� +Ψ0, Ψ′m1, ℓm2+1(A, Ψ′m2) +� +. +(16.32) +In order to proceed, one needs to perform a genus expansion of the various spaces: +this allows to solve recursively for all Bg,n starting from B0,3. One can then build |δΨ′⟩ +recursively, which provides the field redefinition. Indeed, the RHS of this equation contains +only Bg′,n′ for g′ < g or n′ < n and the equation for B0,3 contains no Bg,n in the RHS. +It should be noted that the field redefinition is not unique, but there is the freedom of +performing (infinite-dimensional) gauge transformations. Finding an obstruction to solve +these equations mean that the field redefinition does not exist, and thus that the theory is +not background independent +The form of the equation +∂B0,3 = V0,3 − V′ +0,3 +(16.33) +suggests to use homology theory. The interpretation of B0,3 is that it is a space interpolating +between V0,3 and V′ +0,3. A preliminary step is to check that there is no obstruction: since the +LHS is already a boundary one has ∂2B0,3 = 0 and one should check that ∂(RHS) = 0 as +well. It can be shown that it is indeed true. It was proved in [226] that this equation admits +a solution and that the equations for higher g and n can all be solved. Hence, there exists +a field redefinition and SFT is background independent. +16.7 +Suggested readings +• Proof of the background independence under marginal deformations [226, 232, 233] +(see also [209–211] for earlier results laying foundations for the complete proof). +• L∞ perspective [169, sec. 4] (see also [168, 167, sec. III.B]. +• Connection on the space of CFTs [30, 198, 199, 210, 238]. +244 + +Chapter 17 +Superstring +Abstract +Superstring theory is generally the starting point for physical model building. +It has indeed several advantages over the bosonic string, most importantly, the removal +of the tachyon and the inclusion of fermions in the spectrum. The goal of this chapter is +to introduce the most important concepts needed to generalize the bosonic string to the +superstring, both for off-shell amplitudes and string field theory. We refer to the review [42] +for more details. +17.1 +Worldsheet superstring theory +There are five different superstring theories with spacetime supersymmetry: the types I, IIA +and IIB, and the E8 × E8 and SO(32) heterotic models. +In the Ramond–Neveu–Schwarz formalism (RNS), the left- and right-moving sectors of +the superstring worldsheet are described by a two-dimensional super-conformal field theory +(SCFT), possibly with different numbers of supersymmetries. The prototypical example is +the heterotic string with N = (1, 0) and we will focus on this case: only the left-moving +sector is supersymmetric, while the right-moving is given by the same bosonic theory as in +the other chapters. Up to minor modifications, the type II theory follows by duplicating the +formulas of the left-moving sector to the right-moving one. +17.1.1 +Heterotic worldsheet +The ghost super-CFT is characterized by anti-commuting ghosts (b, c) (left-moving) and +(¯b, ¯c) (right-moving) with central charge c = (−26, −26), associated to diffeomorphisms, and +by commuting ghosts (β, γ) with central charge c = (11, 0), associated to local supersym- +metry. As a consequence the matter SCFT must have a central charge c = (15, 26). If +spacetime has D non-compact dimensions, then the matter CFT is made of: +• a free theory of D scalars Xµ and D left-moving fermions ψµ (µ = 0, . . . , D − 1) such +that cfree = 3D/2 and ¯cfree = D; +• an internal theory with cint = 15 − 3D/2 and ¯cint = 26 − D. +The critical dimension is reached when cint = 0 which corresponds to D = 10. +The diffeomorphisms are generated by the energy–momentum tensor T(z); correspond- +ingly, supersymmetry is generated by its super-partner G(z) (sometimes also denoted by +245 + +TF ). The OPEs of the algebra formed by T(z) and G(z) is: +T(z)T(w) ∼ +c/2 +(z − w)4 + +2T(w) +(z − w)2 + ∂T(w) +z − w , +(17.1a) +G(z)G(w) ∼ +2c/3 +(z − w)3 + 2T(w) +(z − w), +(17.1b) +T(z)G(w) ∼ 3 +2 +G(w) +(z − w)2 + ∂G(w) +(z − w). +(17.1c) +The superconformal ghosts form a first-order system (see Section 7.2) with ϵ = −1 and +λ = 3/2. Hence, they have conformal weights +h(β) = +�3 +2, 0 +� +, +h(γ) = +� +−1 +2, 0 +� +(17.2) +and OPEs +γ(z)β(w) ∼ +1 +z − w, +β(z)γ(w) ∼ − +1 +z − w. +(17.3) +The expressions of the ghost energy–momentum tensors are +T gh = −2b ∂c + c∂b, +T βγ = 3 +2 β∂γ + 1 +2 γ ∂β. +(17.4) +The ghost numbers of the different fields are +Ngh(b) = Ngh(β) = −1, +Ngh(c) = Ngh(γ) = 1. +(17.5) +The worldsheet scalars satisfy periodic boundary conditions. On the other hand, fermions +can satisfy anti-periodic or periodic conditions: this leads to two different sectors, called +Neveu–Schwarz (NS) and Ramond (R) respectively. +βγ system +The βγ system can be bosonized as +γ = η eφ, +β = ∂ξ e−φ, +(17.6) +where (ξ, η) are fermions with conformal weights 0 and 1 (this is a first-order system with +ϵ = 1 and λ = 1), and φ is a scalar field with a background charge (Coulomb gas). This +provides an alternative representation of the delta functions: +δ(γ) = e−φ, +δ(β) = eφ. +(17.7) +Introducing these operators is necessary to properly define the path integral with bosonic +zero-modes. They play the same role as the zero-modes insertions for fermionic fields needed +to obtain a finite result (see also Appendix C.1.3): +� +dc0 = 0 +=⇒ +� +dc0 c0 = 1, +(17.8) +because c0 = δ(c0). For a bosonic path integral, one needs a delta function: +� +dγ0 = ∞ +=⇒ +� +dγ0 δ(γ0) = 1. +(17.9) +246 + +By definition of the bosonization, one has: +T βγ = T ηξ + T φ, +(17.10) +where +T ηξ = −η ∂ξ, +T φ = −1 +2 (∂φ)2 − ∂2φ. +(17.11) +The OPE between the new fields are: +ξ(z)η(w) ∼ +1 +z − w, +eq1φ(z)eq2φ(w) ∼ e(q1+q2)φ(w) +(z − w)q1q2 , +∂φ(z)∂φ(w) ∼ − +1 +(z − w)2 . +(17.12) +The simplest attribution of ghost numbers to the new fields is: +Ngh(η) = 1, +Ngh(ξ) = −1, +Ngh(φ) = 0. +(17.13) +To the scalar field φ is associated another U(1) symmetry whose quantum number is +called the picture number Npic. The picture number of η and ξ are assigned1 such that β +and γ have Npic = 0: +Npic(eqφ) = q, +Npic(ξ) = 1, +Npic(η) = −1. +(17.14) +Because of the background charge, this symmetry is anomalous and correlation functions +are non-vanishing if the total picture number (equivalently the number of φ zero-modes) is: +Npic = 2(g − 1) = −χg. +(17.15) +For the same reason, the vertex operators eqφ are the only primary operators: +h(eqφ) = −q +2(q + 2), +(17.16) +and the Grassmann parity of these operators is (−1)q. Special values are +h(eφ) = 3 +2, +h(e−φ) = 1 +2. +(17.17) +The superstring theory features a Z2 symmetry called the GSO symmetry. All fields are +taken to be GSO even, except β and γ which are GSO odd and eqφ whose parity is (−1)q. +Physical states in the NS sector are restricted to be GSO even: it is required to remove +the tachyon of the spectrum and to get a spacetime with supersymmetry. In type II, the +Ramond sector can be projected in two different ways, leading to the type IIA and type IIB +theories. +The components of the BRST current are: +jB = c(T m + T βγ) + γG + bc∂c − 1 +4 γ2b, +(17.18a) +¯ȷB = ¯c ¯T m + ¯b¯c¯∂¯c. +(17.18b) +From there, it is useful to define the picture changing operator (PCO): +X(z) = {QB, ξ(z)} = c∂ξ + eφG − 1 +4 ∂η e2φ b − 1 +4 ∂(η e2φb), +(17.19) +which is a weight-(0, 0) primary operator which carries a unit picture number. It is obviously +BRST exact. This operator will be necessary to saturate the picture number condition: the +1Any linear combination of both U(1) could have been used. The one given here is conventional, but +also the most convenient. +247 + +naive insertion of eφ ∼ δ(β) breaks the BRST invariance. The PCO zero-mode is obtained +from the contour integral: +X0 = +1 +2πi +� dz +z X(z). +(17.20) +It can be interpreted as delocalizing a PCO insertion from a point to a circle, which decreases +the risk of divergence. +17.1.2 +Hilbert spaces +The description in terms of the (η, ξ, φ) fields leads to a subtlety: the bosonization involves +only the derivative ∂ξ and not the field ξ itself, meaning that the zero-mode ξ0 is absent from +the original Hilbert space defined from (β, γ). In the bosonized language, the Hilbert space +without the ξ zero-mode is called the small Hilbert space and is made of state annihilated +by η0 (the η zero-mode) +Hsmall = +� +|ψ⟩ | η0 |ψ⟩ = 0 +� +. +(17.21) +Removing this condition leads to the large Hilbert space:2 +Hsmall = Hlarge ∩ ker η0. +(17.22) +A state in Hsmall contains ξ with at least one derivative acting on it. +A correlation function defined in terms of the (η, ξ, φ) system is in the large Hilbert space +and will vanish since there is no ξ factor to absorb the zero-mode of the path integral. As +a consequence, correlation functions (and the inner product) are defined with a ξ0 insertion +(by convention at the extreme left) or, equivalently, ξ(z). The position does not matter since +only the zero-mode contribution survives, and the correlation function is independent of z. +Sometimes it is more convenient to work in the large Hilbert space and to restrict later to +the small Hilbert space. +The SL(2, C) invariant vacuum is normalized as +⟨k| c−1¯c−1c0¯c0c1¯c1 e−2φ(z) |k′⟩ = (2π)Dδ(D)(k + k′). +(17.23) +Remark 17.1 (Normalization in type II) In type II theory, the SL(2, C) is normalised +as: +⟨k| c−1¯c−1c0¯c0c1¯c1 e−2φ(z)e− ¯φ( ¯ +w) |k′⟩ = −(2π)Dδ(D)(k + k′). +(17.24) +The sign difference allows to avoid sign differences between type II and heterotic string +theories in most formulas [42]. +The Hilbert space of GSO even states satisfying the b− +0 = 0 and L− +0 = 0 conditions is +denoted by HT (ghost and picture numbers are arbitrary). This Hilbert space is the direct +sum of the NS and R Hilbert spaces: +HT = HNS ⊕ HR. +(17.25) +The subspace of states with picture number Npic = n is written Hn. The picture number +of NS and R states are respectively integer and half-integer. Two special subspaces of HT +play a distinguished role: +� +HT = H−1 ⊕ H−1/2, +� +HT = H−1 ⊕ H−3/2. +(17.26) +2The relation between the small and large Hilbert spaces is similar to the one between the H and +H0 = b0H Hilbert space from the open string since the (b, c) and (η, ξ) are both fermionic first-order +systems. +248 + +To understand this, consider the vacuum |p⟩ of the φ field with picture number p: +|p⟩ = epφ(0) |0⟩ . +(17.27) +Then, acting on the vacuum with the βn and γn modes implies +∀n ≥ −p − 1 +2 : +βn |p⟩ = 0, +∀n ≥ p + 3 +2 : +γn |p⟩ = 0. +(17.28) +For p = −1, all positive modes (starting with n = 1/2) annihilate the vacuum in the NS +sector. This is a positive asset because positive modes which do not annihilate the vacuum +can create states with arbitrary negative energy (since it is bosonic).3 For p = −1/2 or +p = −3/2, the vacuum is annihilated by all positive modes, but not by one of the zero-mode +γ0 or β0. Nonetheless, one can show that the propagator in the R sector allows to propagate +only a finite number of states if one chooses H−1/2; the role of H−3/2 will become apparent +when discussing how to build the superstring field theory. +Basis states are introduced as in the bosonic case: +� +HT = Span{|φr⟩}, +� +HT = Span{|φc +r⟩} +(17.29) +such that +⟨φc +r|φs⟩ = δrs. +(17.30) +The completeness relations are +1 = +� +r +|φr⟩⟨φc +r| +(17.31) +� +HT , and +1 = +� +r +(−1)|φr| |φc +r⟩⟨φr| +(17.32) +on � +HT . +Finally, the operator G is defined as: +G = +� +1 +NS sector, +X0 +R sector. +(17.33) +Note the following properties +[G, L± +0 ] = [G, b± +0 ] = [G, QB] = 0. +(17.34) +It will be appear in the propagator and kinetic term of the superstring field theory. +17.2 +Off-shell superstring amplitudes +In this section, we are going to build the scattering amplitudes. +The procedure is very +similar to the bosonic case, except for the PCO insertions and of the Ramond sector. For +this reason, we will simply state the result and motivate the modifications with respect to +the bosonic case. +3This is not a problem on-shell since the BRST cohomology is independent of the picture number. +However, this matters off-shell since such states would propagate in loops and make the theory inconsistent. +249 + +17.2.1 +Amplitudes +External states can be either NS or R: the Riemann surface corresponding to the g-loop +scattering of m external NS states and n external R states is denoted by Σg,m,n. R states +must come in pairs because they correspond to fermions. As in the bosonic case, the amp- +litude is written as the integration of an appropriate p-form Ω(g,m,n) +p +over the moduli space +Mg,m,n (or, more precisely, of a section of a fibre bundle with this moduli space as a basis). +From the geometric point of view, nothing distinguishes the punctures and thus: +Mg,m,n := dim Mg,m,n = 6g − 6 + 2m + 2n. +(17.35) +The form ΩMg,m,n is defined as a SCFT correlation function of the physical vertex operators +together with ghost and PCO insertions. +Remark 17.2 A simple way to avoid making errors with signs is to multiply every Grass- +mann odd external state with a Grassmann odd number. These can be removed at the end +to read the sign. +The two conditions from the U(1) anomalies on the scattering amplitude are: +Ngh = 6 − 6g, +Npic = 2g − 2. +(17.36) +Given an amplitude with m NS states V NS +i +∈ H−1 and n R states V R +j +∈ H−1/2, the above +picture number can be reached by introducing a certain number of PCO X(yA): +npco := 2g − 2 + m + n +2 . +(17.37) +These PCO are inserted at various positions: while the amplitude does not depend on these +locations on-shell, off-shell it will (because the vertex operators are not BRST invariant). +The choices of PCO locations are arbitrary except for several consistency conditions: +1. avoid spurious poles (Section 17.2.3); +2. consistent with factorization (each component of the surface in the degeneration limits +must saturate the picture number condition). +This parallels the discussion of the choices of local coordinates: as a consequence, the +natural object is a fibre bundle �Pg,m,n with the local coordinate choices (up to global phase +rotations) and the PCO locations as fibre, and the moduli space Mg,m,n as base. Forgetting +about the PCO locations leads to a fibre bundle �Pg,m,n which is a generalization of the +one found in the bosonic case. The coordinate system of the fibre bundle presented in the +bosonic case is extended by including the PCO locations {yA}. +With these information, the amplitude can be written as: +Ag,m,n(V NS +i +, V R +j ) = +� +Sg,m,n +ΩMg,m,n(V NS +i +, V R +j ), +(17.38a) +where +ΩMg,m,n = (−2πi)−Mc +g,m,n +�Mg,m,n +� +λ=1 +Bλ dtλ +npco +� +A=1 +X(yA) +m +� +i=1 +V NS +i +n +� +j=1 +V R +j +� +Σg,n +. +(17.38b) +where Sg,m,n is a Mg,m,n-dimensional section of �Pg,m,n parametrized by coordinates tλ. The +1-form B corresponds to a generalization of the bosonic 1-form. +It has ghost number 1 +250 + +and includes a correction to compensate the variation of the PCO locations in terms of the +moduli parameters: +Bλ = +� +α +� +Cα +dσα +2πi b(σα) ∂Fα +∂tλ +� +F −1 +α (σα) +� ++ +� +α +� +Cα +d¯σα +2πi +¯b(¯σα) ∂ ¯Fα +∂tλ +� ¯F −1 +α (¯σα) +� +− +� +A +1 +X(yA) +∂yA +∂tλ +∂ξ(yA). +(17.39) +The last factor amounts to consider the combination +X(yA) − ∂ξ(yA) dyA +(17.40) +for each PCO insertion:4 the correction is necessary to ensure that the BRST identity +(13.46) holds. This can be understood as follows: the derivative acting on the PCO gives a +term dX(z) = ∂X(z)dz which must be cancelled. This is achieved by the second term since +{QB, ∂ξ(z)} = ∂X(z). +Remark 17.3 While it is sufficient to work with Mg,n for on-shell bosonic amplitudes, +on-shell superstring amplitudes are naturally expressed in �Pg,m,n (with the local coordinate +removed) since the positions of the PCO must be specified even on-shell. +Remark 17.4 (Amplitudes on the supermoduli space) Following Polyakov’s appro- +ach from Chapters 2 and 3 to the superstring would lead to replace the moduli space by the +supermoduli space. The latter includes Grassmann-odd moduli parameters in addition to the +moduli parameters from Mg,m+n (in the same way the superspace includes odd coordinates +θ along with spacetime coordinates x). The natural question is whether it is possible to split +the integration over the even and odd moduli, and to integrate over the latter such that only +an integral over Mg,m+n remains. In view of (17.38a), the answer seems positive. However, +this is incorrect: it was proven in [58] that there is no global holomorphic projection of the +supermoduli space to the moduli space. +This is related to the problem of spurious poles +described below. But, this does not prevent to do it locally: in that case, implementing the +procedure carefully should give the rules of vertical integration [73, 215, 230]. +17.2.2 +Factorization +The plumbing fixture of two Riemann surfaces Σg1,m1,n1 and Σg2,m2,n2 can be performed in +two different ways since two NS or two R punctures can be glued. +If two NS punctures are glued, the resulting Riemann surface is Σ(NS) +g1+g2,m1+m2−2,n1+n2. +The number of PCO inherited from the two original surfaces is +n(1) +pco + n(2) +pco = 2(g1 + g2) − 2 + (m1 + m2 − 2) + n1 + n2 +2 += n(NS) +pco , +(17.41) +which is the required number for a non-vanishing amplitude. As a consequence, the propag- +ator is the same as in the bosonic case: +∆NS = b+ +0 b− +0 +1 +L0 + ¯L0 +δ(L− +0 ). +(17.42) +If two R punctures are glued, the numbers of PCO do not match by one unit: +n(1) +pco + n(2) +pco = 2(g1 + g2) − 2 + (m1 + m2) + n1 + n2 − 2 +2 +− 1 = n(R) +pco − 1. +(17.43) +4The sum is formal since it is composed of 0- and 1-forms. +251 + +This means that an additional PCO must be inserted in the plumbing fixture procedure: +the natural place for it is in the propagator since this is the only way to keep both vertices +symmetric as required for a field theory interpretation. Another way to see the need of +this modification is to study the propagator (17.42) for Ramond states: since Ramond +states carry a picture number −1/2, the conjugate states have Npic = −3/2 and thus the +propagator has a total picture number −3 instead of −2 (the propagator graph is equivalent +to a sphere). Then, to avoid localizing the PCO at a point of the propagator, one inserts +the zero-mode which corresponds to smear the PCO: +∆R = b+ +0 b− +0 +X0 +L0 + ¯L0 +δ(L− +0 ). +(17.44) +Delocalizing the PCO amounts to average the amplitude over an infinite number of points +(i.e. to consider a generalized section): this is necessary to preserve the L− +0 eigenvalue since +X0 is rotationally invariant while X(z) is not. +Note that the zero-mode can be written +equivalently as a contour integral around one of the two glued punctures: +X0 = +1 +2πi +� dw(1) +n +w(1) +n +X +� +w(1) +n +� += +1 +2πi +� dw(2) +n +w(2) +n +X +� +w(2) +n +� +. +(17.45) +The equality of both expressions holds because X(z) has conformal weight 0. +Using the operator G (17.33), the propagator can be written generically as +∆ = b+ +0 b− +0 +G +L0 + ¯L0 +δ(L− +0 ). +(17.46) +Remark 17.5 (Propagators) NS and R states correspond respectively to bosonic and fer- +mionic fields: the operators L+ +0 and X0 can be interpreted as the (massive) Laplacian and +Dirac operators, such that both propagators can be written +∆NS ∼ +1 +k2 + m2 , +∆R ∼ i/∂ + m +k2 + m2 . +(17.47) +To motivate the identification of X0 with the Dirac operator, remember that X(z) contains +a term eφ(z)G(z) (this is the only term which contributes on-shell), where G(z) in turn +contains ψµ∂Xµ. But, the zero-modes of ψµ and ∂Xµ correspond respectively to the gamma +matrix γµ and momentum kµ when acting on a state. +The PCO zero-mode insertion inside the propagator has another virtue. It was noted +previously that states with Npic = −3/2 are infinitely degenerate since one can apply β0 +an arbitrary number of time. These states have large negative ghost numbers. Considering +a loop amplitude, all these states would appear in the sum over the states and lead to a +divergence. The problem is present only for loops because the ghost number is not fixed: in +a tree propagator, the ghost number is fixed and only a finite number of β0 can be applied. +But, the PCO insertion turns these states into Npic = −1/2 states. In this picture number, +one cannot create an arbitrarily large negative ghost number since γn +0 can only increase the +ghost number. +17.2.3 +Spurious poles +A spurious pole corresponds to a singularity of the amplitude which cannot be interpreted +as the degeneration limit of Riemann surfaces. As a consequence, they do not correspond to +infrared divergences and don’t have any physical meaning; they must be avoided in order to +define a consistent theory. To achieve this, the section Sg,m,n must be chosen such that it +252 + +avoids all spurious poles. However, while it is always possible to avoid these poles locally, it +is not possible globally (this is related to the results from [58]). Poles can be avoided using +vertical integration: two methods have been proposed, in the small (Sen–Witten) [215, 230] +and large (Erler–Konopka) [73] Hilbert spaces respectively. Before describing the essence of +both approaches, we review the origin of spurious poles. +Origin +Spurious poles arise in three different ways: +• two PCOs collide; +• one PCO and one matter vertex collide; +�� other singularities of the correlation functions. +The last source is the less intuitive one and we focus on it. +A general correlation function of (η, ξ, φ) on the torus5 (satisfying the ghost number +condition) reads +C(xi, yj, zq) = +�n+1 +� +i=1 +ξ(xi) +n +� +j=1 +η(yj) +m +� +k=1 +eqkφ(zk) +� += +n� +j′=1 +ϑδ +� +− yj′ + � +i +xi − � +j +yj + � +k +qkzk +� +n+1 +� +i′=1 +ϑδ +� +− xi′ + � +i +xi − � +j +yj + � +k +qkzk +� × +� +i 0 is a number. There is an implicit sum over the momentum +indices. +The terms quadratic in the momenta inside the exponential arise from two sources: +• The correlation functions of the vertex operators ⟨� +i eiki·X(zi)⟩ is proportional to +e−ki·kjG(zi,zj), where G is the Green function. +Additional factors like ∂X contrib- +ute to the polynomial Pα1,...,αn. +• It is possible to add stubs to the vertices. The effect is to multiply each leg by a factor +e−λ(k2 +i +m2 +i ) with λ > 0 (we take λ to be the same for all vertices for simplicity). The +first term of the exponential contributes to the diagonal of the matrix gij. By taking +λ sufficiently large, one can enforce that all eigenvalues are positive. +Finally, the exponential term with the masses m2 +α ensures that the sum over all intermediate +states converge despite an infinite number of states. Indeed, the number of states of mass +261 + +mα grows as ecmα, which is dominated by e−λm2 +α for sufficiently large λ. Hence, the addition +of stubs make explicit the absence of divergences in SFT.1 +The vertices have no singularity for ki ∈ C finite. As the energy becomes infinite |k0 +i | → +∞, they behave as: +lim +k0→±i∞ V (n) = 0, +lim +k0→±∞ V (n) = ∞. +(18.5) +The first property is responsible for the soft UV behaviour of string theory in Euclidean +signature, while the second prevents from performing the Wick rotation (indeed, the pole +at infinity implies that the arcs closing the contour contribute). +The g-loop n-point amputated Green functions are sums of Feynman diagrams, each of +the form: +Fg,n(p1, . . . , pn) ∼ +� +dT +� +s +dDℓs e−Grs(T ) ℓr·ℓs−2Hri(T ) ℓr·pi−Fij(T ) pi·pj +× +� +a +1 +k2a + m2a +P(pi, ℓr; T), +(18.6) +where {pi} are the external momenta, {ℓr} the loop momenta and {ki} the internal mo- +menta, with the latter given by a linear combination of the other. Moreover, T denotes the +dependence in the moduli parameters of all the internal vertices, and P is a polynomial in +(pi, ℓr). The matrix Grs is positive definite, which implies that: +• integrations over spatial loop momenta ℓr converge; +• integrations over loop energies ℓ0 +r diverge. +As a consequence, the Feynman diagrams in Lorentzian signature are ill-defined: we will +explain in the next section how to fix this problem. +18.2 +Generalized Wick rotation +We have seen that loop integrals in Lorentzian signature are divergent because of the large +energy behaviour of the interactions. But, this is not different from the usual QFT, where the +loop integrals are also ill-defined in Lorentzian signature. Indeed, poles of the propagators +sit on the real axis and also give divergent loop integrals (note that the same problem arise +also here). In that case, the strategy is to define the Feynman diagrams in Euclidean space +and to perform a Wick rotation: the latter matches the expressions in Lorentzian signature +up to the iε-prescription. The goal of the latter is to move slightly the poles away from the +real axis. +Example 18.1 – Scalar field +Consider a scalar field of mass m with a quartic interaction. The 1-loop 4-point Feyn- +man diagram is given in Figure 18.1. The external momenta are pi, i = 1, . . . , 4. There +are one loop momentum ℓ and two internal momenta k1 = ℓ and k2 = p − ℓ, where +p = p1 + p2. The poles in the loop energy ℓ0 are located at: +p± = ± +� +ℓ2 + m2, +q± = p0 ± +� +(p − ℓ)2 + m2. +(18.7) +The graph is first defined in Euclidean signature, where the external and loop en- +ergies are pure imaginary, p0 +i , ℓ0 ∈ iR. The poles are shown in Figure 18.2. Then, +the external momenta are analytically continued to real values, p0 +i ∈ R. At the same +1Remember that λ is not a physical parameter and disappears on-shell. This means that the cancellation +of the divergences is independent of λ and must always happen on-shell. +262 + +time, the integration contour is also analytically continued thanks to the Wick rotation +(Figure 18.3). The contour is closed with arcs, but they don’t contribute since there is +no poles in the upper-right and lower-left quadrants, and no poles at infinity. However, +one cannot continue the contour such that ℓ0 ∈ R because of the poles on the real axis. +The Wick rotation is possible for ℓ0 in the upper-right quadrant, Re ℓ0 ≥ 0, Im ℓ0 > 0, +which leads to the iε-prescription ℓ0 ∈ R + iε. +Figure 18.1: 1-loop 4-point function for a scalar field theory. +Figure 18.2: Integration contour for external Euclidean momenta. +Since the Feynman diagram (18.6) is not defined in Lorentzian signature because of the +poles at ℓ0 +r → ±∞, it is also necessary to start with Euclidean momenta. However, the +same behaviour at infinity prevents from using the Wick rotation since the contribution +from the arcs does not vanish. It is then necessary to find another prescription for defining +the Feynman diagrams in SFT starting from the Euclidean Green functions. This is given +by the following generalized Wick rotation (Pius–Sen [187]): +1. Define the Green functions for Euclidean internal and external momenta. +2. Perform an analytic continuation of the external energies and of the integration contour +such that: +• keep poles on the same side; +• keep the contour ends fixed at ±i∞. +One can show [187] that the Green functions are analytic in the upper-right quadrant Im p0 +a > +0, Re p0 +a ≥ 0, for pa ∈ R, p0 +a. Moreover, the result is independent of the contour chosen as +long as it satisfies the conditions described above. In fact, this generalized Wick rotation is +263 + +Figure 18.3: Integration contour for external Lorentzian momenta after Wick rotation (reg- +ular vertices). +valid even for normal QFT, which raises interesting questions. For example, it seems that +the internal and external set of states have no intersection, which can be puzzling when +trying to interpret the Cutkosky rules. Nonetheless, everything works as expected. +Remark 18.1 (Timelike Liouville theory) It has been shown in [13] that this general- +ized Wick rotation is also the correct way for defining the timelike Liouville theory. +The fact that the amplitude is analytic only when the imaginary parts of the momenta +are not zero, Im p0 +a > 0, is equivalent to the usual iε-prescription for QFT. Moreover, it has +been shown [222] to be equivalent to the moduli space iε-prescription from [257]. Then, it +has also been used to prove several important properties of string theory shared by local +QFTs: Cutkosky rules [187, 188], unitarity [220, 221], analyticity in a subset of the primitive +domain and crossing symmetry [43]. Finally, general soft theorems for string theory (and, in +fact, any theory of quantum gravity) have been proven in [34, 150, 223, 224]. All together, +these properties establish string theory as a very strong candidate for a consistent theory +of everything. +The next main question is how to obtain an expression of SFT which is +amenable to explicit computations. This will certainly require to understand even better +the deep structure of SFT, a goal which this book will hopefully help the reader to achieve. +18.3 +Suggested readings +• SFT momentum space action [42, 187, 225]. +• Consistency properties of string theory [42]: +– generalized Wick rotation, Cutkosky rules and unitarity [187, 188, 219–222]. +– analyticity and crossing symmetry [43]. +– soft theorems [34, 150, 223, 224]. +264 + +(a) +(b) +(c) +Figure 18.4: Integration contour after analytic continuation to external Lorentzian momenta. +Depending on the values of the external momenta, different cases can happen. +265 + +Appendix A +Conventions +Most of the book uses natural units where c = ℏ = 1, but the string length ℓs (or Regge +slope α′) are kept. +A bar is used to denote both complex conjugation and the anti-holomorphic operators. +The symbol := (resp. =:) means that the LHS (RHS) is defined by the expression in the +RHS (LHS). +A.1 +Coordinates +The number of spacetime (target-space) dimensions is denoted by D = d + 1, where d is +the number of spatial dimensions. The corresponding spacetime and spatial coordinates are +written with Greek and Latin indices: +xµ = (x0, xi), +µ = 0, . . . , D − 1 = d +i = 1, . . . , d +(A.1) +When time is singled out, one writes x0 = t in Lorentzian signature and x0 = tE in Euclidean +signature (or x0 = τ when there is no ambiguity with the worldsheet time). +A p-brane is a (p+1)-dimensional object whose worldvolume is parametrized by coordin- +ates: +σa = (σ0, σα), +a = 0, . . . , p − 1, +α = 1, . . . , p. +(A.2) +The time coordinate can also be singled out as σ0 = τM in Lorentzian signature and as +σ0 = τ in Euclidean signature. For the string, the index α is omitted since it takes only one +value. +The Lorentzian signature is taken to be mostly plus and the flat Minkowski metric reads +ηµν = diag(−1, 1, . . . , 1 +� �� � +d +). +(A.3) +The flat Euclidean metric is +δµν = diag(1, . . . , 1 +� �� � +D +). +(A.4) +Similar notations hold for the worldvolume metrics ηab and δab. The Levi–Civita (completely +antisymmetric) tensor is normalized by +ϵ01 = −ϵ01 = 1. +(A.5) +Wick rotation from Lorentzian time t to Euclidean time τ (either worldsheet or target +spacetime) is defined by +t = −iτ. +(A.6) +266 + +Accordingly, contravariant (covariant) vector transforms with the same (opposite) factor: +V 0 +M = −iV 0 +E, +VM,0 = iVE,0. +(A.7) +Most computations are performed with both spacetime and worldsheet Euclidean signatures. +Expressions are Wick rotated when needed. +Light-cone coordinates are defined by +x± = x0 ± x1. +(A.8) +A function depending only on x+ (x−) is said to be left-moving (right-moving) by ana- +logy with the displacement of a wave. Under analytic continuation, the left-moving (right- +moving) coordinate is mapped to the holomorphic1 (anti-holomorphic) coordinate z (¯z). In +chiral theories, the left-moving value is written first. +The worldsheet coordinates (τ, σ) on the cylinder are defined by +τ ∈ R, +σ ∈ [0, L), +σ ∼ σ + L, +(A.9) +where typically L = 2π. The integration over the spatial coordinate is normalized such that +the perimeter of spatial slice is normalized to 1 if L = 2π: +L = 1 +2π +� L +0 +dσ = L +2π . +(A.10) +This implies that 2d action, conserved charges, etc. are divided by an extra factor of 2π. +The coordinates can be written in terms of complex coordinates +w = τ + iσ, +¯w = τ − iσ +(A.11) +such that the flat metric is +ds2 = dτ 2 + dσ2 = dwd ¯w. +(A.12) +Under Wick rotation, the complex coordinates are mapped to light-cone coordinates as +follows: +w = iσ+, +¯w = iσ−. +(A.13) +The cylinder can be mapped to the complex plane through +z = e2πw/L, +¯z = e2π ¯ +w/L. +(A.14) +The definition of the Levi–Civita tensor includes the √g factor, such that +ϵz¯z = i +2, +ϵz¯z = −2i +(A.15) +on the complex plane with flat metric. +1The terms of holomorphic is simply used to indicate that the object depends only on z, but not on ¯z. +Typically, the objects have singularities and are really meromorphic in z. +2In fact, the terms of “left”- and “right”-moving are interchanged in [193, p. 34] to get agreement with +the literature. But, it means that the spatial axis orientation is reversed. +Moreover, concerning [54], the first definition agrees with (6.1) but not with (6.53) since the definition of +ξ (our w) is modified in-between. This explains why the definitions of left- and right-moving [54, p. 161] do +not agree with the one given in the table. +267 + +refs +cylinder +plane +light-cone +left-moving +here, Di Francesco et al. +[25, 54, 124, 126, 128, 212, 252, 265] +w = τ + iσ +z = ew +w = iσ+, ¯w = iσ− +holomorphic +Blumenhagen et al. +[14, 24, 123, 205] +w = τ − iσ +z = ew +w = iσ−, ¯w = iσ+ +anti-holomorphic +Polchinski [176, 193, 246] +w = σ + iτ +z = e−iw +w = −σ−, ¯w = σ+ +anti-holomorphic2 +Table A.1: Conventions for the coordinates.The notations are the following (they can slightly +vary depending on the references): the Euclidean time is obtained by the analytic continu- +ation τ = it (denoted also by τ = σ0 = σ2) the spatial direction is σ = σ1, and the light-cone +coordinates are σ± = t ± σ. +A.2 +Operators +Commutators and anti-commutators are denoted by +[A, B] := [A, B]− = AB − BA, +{A, B} := [A, B]+ = AB + BA. +(A.16) +The Grassmann parity of a field A is denoted by |A| +|A| = +� ++1 +Grassmann odd, +0 +Grassmann even. +(A.17) +Two (anti-)commuting operators satisfy +AB = (−1)|A||B|BA. +(A.18) +A.3 +QFT +Energy is defined as the first component of the momentum vector +pµ := (E, pi). +(A.19) +The following notations are used to denote the number of supersymmetries: +(NL, NR), +N = NL + NR, +(A.20) +where NL and NR are the numbers of left- and right-chirality supersymmetries. The last +form is used when it is not important to know the chirality of the supercharges. +The variation of a field φ(x) is defined by +δφ(x) = φ′(x) − φ(x). +(A.21) +Given an internal symmetry with parameters αa, the Noether current in Lorentzian signature +is given by: +Jµ +a = λ +∂L +∂(∂µφ) +δφ +δαa , +∇µJµ +a = 0, +(A.22) +where L is the Lagrangian which does not include the factor √g for curved spaces and λ is +some normalization.3 The conserved charges Qa associated to the currents Jµ +a for a fixed +spatial slice t = cst are +Qa = 1 +λ +� +Σ +dD−1x +√ +h J0 +a, +(A.23) +3Including the √g would give the current density √gJµ +a . The simple derivative of the latter vanishes +∂µ(√gJµ +a ) = 0 in view of the identity (B.4). +268 + +where Σ is a spatial slice and h is the induced metric. One sets λ = 2π in two dimensions, +otherwise λ = 1. The variation of a field under a transformation generated by Q is +δαaφ(x) = iαa[Qa, φ(x)] +(A.24) +In Euclidean signature, the current and variation are: +Jµ +a = iλ +∂L +∂(∂µφ) +δφ +δαa , +(A.25a) +δαaφ(x) = −αa[Qa, φ(x)]. +(A.25b) +Note that the charge is still given by (A.23). The factor of i in (A.25a) can be understood +as follows.4 +First, the time component J0 +a of the current transforms like time such that +J0 +a → iJ0 +a, which implies that the charge also gets a factor i, Qa → iQa. This explains the +minus sign in (A.25b). Then, one needs to make this consistent with the formula (B.9) for +the charge associated to a general surface. Given a spacelike nµ, the integration measure +includes the time which transforms with a factor of i: one can interpret it as coming from the +spatial components of the current, Ji +a → iJi +a, while working with a Euclidean region. Another +way to understand this factor for the spatial vector is by considering the electromagnetic +case, where J contains a time derivative. +The term “zero-mode” has two (related) meanings: +1. given an operator D acting on a space of fields ψ(z), zero-modes ψ0,i(z) of the operator +are all fields with zero eigenvalue Dψ0,i(z) = 0, i = 1, . . . , dim ker D +2. the zero-mode of a field expansion ψ = � +n ψnz−n−h is the mode ψ0 for n = 0: on the +cylinder, it corresponds to the constant term of the Fourier expansion on the cylinder +(hence, a zero-mode of ∂z according to the previous definition) +A prime indicates that the zero-modes are excluded. For example, det′ D is the product of +non-zero eigenvalues, φ′ is a field without zero-mode and d′φ the corresponding integration +measure. +A.4 +Curved space and gravity +The covariant derivative is defined by +∇µ = ∂µ + Γµ +(A.26) +where Γµ is the connection. For example, one has for a vector field +∇µAν = ∂µAν + Γ +ν +µρ Aρ. +(A.27) +The negative-definite Laplacian (or Laplace–Beltrami operator) is defined by +∆ = gµν∇µ∇ν = +1 +√g ∇µ +�√ggµν∇ν). +(A.28) +Note that ∇µ does not contain the Christoffel symbol for the index ν because of the identity +(B.4) (but it contains a connection for any other index of the field). For a scalar field, both +derivatives become simple derivatives. +The energy–momentum tensor is defined by +Tµν = − 2λ +√g +δS +δgµν , +(A.29) +where λ = 2π for D = 2 and λ = 1 otherwise. +4We stress that these formulas and arguments do not apply to the energy–momentum tensor. +269 + +A.5 +List of symbols +General: +• D: number of non-compact spacetime dimensions +• g: loop order for a scattering amplitude +• n: number of external closed string states +• xµ: spacetime non-compact coordinates +• σa = (t, σ): worldsheet coordinates +• gs: closed string coupling +• Zg = Ag,0: genus-g vacuum amplitude +• Ag,n(k1, . . . , kn)α1,...,αn := Ag,n({ki}){αi}: g-loop n-point scattering amplitude for +states with quantum numbers {ki, αi} (if connected, amputated Green functions for +n ≥ 3) +• Gg,n(k1, . . . , kn)α1,...,αn: g-loop n-point Green function for states with quantum num- +bers {ki, αi} +• T ⊥ +ab: traceless symmetric tensor or traceless component of the tensor Tab +• Ψ: generic (set of) matter field(s) +Hilbert spaces: +• H: generic Hilbert space (in general, Hilbert space of the matter plus ghost CFT) +• H± = H ∩ ker b± +0 +• H0 = H ∩ ker b− +0 ∩ ker b+ +0 +• H(QB): absolute cohomology of the operator QB inside the space H +• H−(QB) = H(QB) ∩ H−: semi-relative cohomology of the operator QB inside the +space H +• H0(QB) = H(QB) ∩ H0: relative cohomology of the operator QB inside the space H +• A: Grassmann parity of the operator or state A +Riemann surfaces: +• g: Riemann surface genus (number of holes / handles) +• n: number of bulk punctures / marked points +• Σg,n: genus-g Riemann surface with n punctures +• Σg = Σg,0: genus-g Riemann surface +• Mg,n: moduli space of genus-g Riemann surfaces with n punctures +• Mg = Mg,0: moduli space of genus-g Riemann surfaces +• Mg,n = dim Mg,n +270 + +• Mc +g,n = dimC Mg,n +• Mg = Mg,0 = dim Mg = dim ker P † +1 +• Mc +g = Mc +g,0 = dimC Mg +• Kg,n: conformal Killing vector group of genus-g Riemann surfaces with n punctures +• Kg = Kg,0 = ker P1: conformal Killing vector group of genus-g Riemann surfaces +• Kg,n = dim Kg,n +• Kc +g,n = dimC Kg,n = dimC ker P1 +• Kg = Kg,0 = dim ker P1 +• Kc +gKc +g,0 = dimC ker P1 +• ψi: real basis of ker P1, CKV +• φi: real basis of ker P † +1 , real quadratic differentials +• (ψK, ¯ψK): complex basis of ker P1, (anti-)holomorphic CKV +• (φI, ¯φI): complex basis of ker P † +1 , (anti-)holomorphic quadratic differentials +• tλ ∈ Mg,n: real moduli of Mg,n +• mΛ ∈ Mg,n: complex moduli of Mg,n +• ti ∈ Mg: real moduli of Mg +• mI ∈ Mg: complex moduli of Mg +• z: coordinate on the Riemann surface +• wi: local coordinates around punctures +• za: local coordinates away from punctures +• fi(wi): transition functions from wi to z +• σα: coordinate system on the left of the contour Cα +• τα: coordinate system on the right of the contour Cα +CFT: +• Vα(k; σa) := Vk,α(σa): matter vertex operator with5 momentum k and quantum num- +bers α inserted at position σa = (z, ¯z) +• Vα(k; σa): unintegrated vertex operator with momentum k and quantum numbers α +inserted at position σa +• Vα(k) = +� +d2σ√g Vα(k; σ): integrated vertex operator +• on-shell (closed bosonic string): Vα(k; σa) = c¯cVα(k; σa) is a (0, 0)-primary, with +Vα(k; σa) a (1, 1)-primary matter operator +5When the momentum and/or quantum numbers are not relevant, we remove them or simply index the +operators by a number. +271 + +• � +O: operator O with zero-modes removed +• O†: Hermitian adjoint +• O‡: Euclidean adjoint +• Ot: BPZ conjugation +• ⟨O1|O2⟩: BPZ inner-product +• ⟨O‡ +1|O2⟩: Hermitian inner-product +• |0⟩: SL(2, C) (conformal) vacuum +• |Ω⟩: energy vacuum (lowest energy state) +• :O: : conformal normal ordering (with respect to SL(2, C) vacuum |0⟩) +• +⋆ +⋆O +⋆ +⋆ : energy normal ordering (with respect to energy vacuum |Ω⟩) +SFT: +• Ψ: closed string field +• {φr} = {φα(k)}: basis of H (or some subspace) +Indices: +• µ = 0, . . . , D − 1: non-compact spacetime dimensions +• a = 0, . . . , p: worldvolume coordinates (p = 1: worldsheet) +• i = 1, . . . , n: external states, local coordinates +• λ = 1, . . . , Mg,n: real moduli of Mg,n +• Λ = 1, . . . , Mc +g,n: complex moduli of Mg,n +• i = 1, . . . , Mg: real moduli of Mg +• I = 1, . . . , Mc +g: complex moduli of Mg +• i = 1, . . . , Kg: real CKV of Kg +• K = 1, . . . , Kc +g: complex CKV of Kg +• r = (k, α): index for basis state of H (or some subspaces), α: non-momentum indices +272 + +Appendix B +Summary of important formulas +This appendix summarizes formulas which appear in the book or which are needed but +assumed to be known to the reader (such as formulas from QFT and general relativity). +B.1 +Complex analysis +The Cauchy–Riemann formula is +� +Cz +dw +2πi +f(w) +(w − z)n = f (n��1)(z) +(n − 1)! , +(B.1) +where f(z) is an holomorphic function. +One has +¯∂ 1 +z = 2π δ(2)(z). +(B.2) +B.2 +QFT, curved spaces and gravity +The Green function G of a differential operator D is defined by +DxG(x, y) = δ(x − y) +√g +− P(x, y), +(B.3) +where P is the projector on the zero-modes of D. +The covariant divergence of a vector can be rewritten in terms of a simple derivative: +∇µvµ = +1 +√g ∂µ(√gvµ). +(B.4) +Under an infinitesimal change of coordinates +δxµ = ξµ, +(B.5) +the metric transforms as +δgµν = Lξgµν = ∇µξν + ∇νξµ. +(B.6) +Stokes’ theorem reads +� +V +dDx ∇µvµ = +� +∂V +dΣµvµ, +dΣµ := ϵ nµ dD−1Σ, +(B.7) +273 + +where V is a spacetime region, S = ∂V its boundary and dD−1Σ the induced integration +measure. The vector nµ normal to S points outward and ϵ := nµnµ = 1 (−1) if S is timelike +(spacelike). If the surface is defined by x0 = cst, then +dD−1Σ = √g dD−1x, +nµ = δ0 +µ. +(B.8) +We can write a generalization of (A.23) for a charge associated to a general surface S: +QS = 1 +λ +� +S +dΣµ Jµ +a . +(B.9) +If the current Jµ +a is conserved, ∇µJµ +a = 0 (no source), Stokes’ theorem (B.7) shows that the +charge vanishes QS = 0 if S is a closed surface and that it is conserved QS1 = −QS2 for +two spacelike surfaces S1 and S2 extending to infinity (if Jµ +a vanishes at infinity) (see [190, +chap. 3, 265, sec. 8.4] for more details). +B.2.1 +Two dimensions +Stokes’ theorem (B.7) on flat space reads +� +d2x ∂µvµ = +� +ϵµν dxνvµ = +� +(v0dσ − v1dτ), +(B.10) +since dΣµ = ϵµνdxν. +The integral of the curvature is a topological invariant +χg;b := 1 +4π +� +d2σ√g R + 1 +2π +� +ds k += 2 − 2g − b, +(B.11) +called the Euler characteristics and where g is the number of holes and b the number of +boundaries. +B.3 +Conformal field theory +In two dimensions, the energy–momentum tensor is defined by +Tab = − 4π +√g +δS +δgab . +(B.12) +B.3.1 +Complex plane +Defining the real coordinates (x, y) from the complex coordinate on the complex plane +z = x + iy, +z = x − iy, +(B.13) +274 + +we have the formulas: +ds2 = dx2 + dy2 = dzd¯z, +gz¯z = 1 +2, +gzz = g¯z¯z = 0, +(B.14a) +ϵz¯z = i +2, +ϵz¯z = −2i, +(B.14b) +∂ := ∂z = 1 +2 (∂x − i∂y), +¯∂ := ∂¯z = 1 +2 (∂x + i∂y), +(B.14c) +V z = V x + iV y, +V ¯z = V x − iV y, +(B.14d) +d2x = dxdy = 1 +2 d2z, +d2z = dzd¯z, +(B.14e) +δ(z) = 1 +2 δ(2)(x), +1 = +� +d2z δ(2)(z) = +� +d2x δ(2)(x), +(B.14f) +� +R +d2z (∂zvz + ∂¯zv¯z) = −i +� +∂R +� +dz v¯z − d¯zvz� += −2i +� +∂R +(vzdz − v¯zd¯z). +(B.14g) +B.3.2 +General properties +A primary holomorphic field φ(z) of weight h transforms as +f ◦ φ(z) = +�df +dz +�h +φ +� +f(z) +� +(B.15) +for any local change of coordinates f. A quasi-primary operator transforms like this only +for f ∈ SL(2, C). Its mode expansion reads +φ(z) = +� +n +φn +zn+h , +φn = +� +C0 +dz +2πi zn+h−1φ(z), +(B.16) +where the integration is counter-clockwise around the origin. +The SL(2, C) vacuum |0⟩ is defined by +∀n ≥ −h + 1 : +φn |0⟩ = 0. +(B.17) +Its BPZ conjugate ⟨0| satisfies: +∀n ≤ h − 1 : +⟨0| φn = 0. +(B.18) +The state–operator correspondence associates a state |φ⟩ to each operator φ(z): +|φ⟩ := φ(0) |0⟩ = φ−h |0⟩ . +(B.19) +The operator corresponding to the vacuum is the identity 1.1 +The Hermitian and BPZ +conjugated states are +⟨φ‡| :=⟨0| I ◦ φ†(0) = lim +z→∞ z2h⟨0| φ†(z), +⟨φ| :=⟨0| I± ◦ φ(0) = (±1)h lim +z→∞ z2h⟨0| φ(z). +(B.20) +The energy–momentum tensor is a quasi-primary operator of weight h = 2 +T(z) = +� +n +Ln +zn+2 . +(B.21) +1Exceptionally, the state |0⟩ and the operator 1 does not have the same symbol. +275 + +The OPE between T and a primary operator h of weight h is +T(z)φ(w) ∼ +h φ(w) +(z − w)2 + ∂φ(w) +z − w . +(B.22) +The OPE of T with itself defines the central charge c +T(z)T(w) ∼ +c/2 +(z − w)4 + +2T(w) +(z − w)2 + ∂T(w) +z − w . +(B.23) +B.3.3 +Hermitian and BPZ conjugations +Both conjugations do not change the ghost number of a state. +Hermitian +The Hermitian conjugate of a general state built from n operators Ai and a complex number +λ is +(λ A1 · · · An |0⟩)† = λ∗ ⟨0| A† +n · · · A† +1. +(B.24) +BPZ +The BPZ conjugate of modes is +φt +n = (I± ◦ φ)n = (−1)h(±1)nφ−n, +(B.25) +where I±(z) = ±1/z. The plus sign is usually used for the closed string, and the minus sign +for the open string. Given a general state built from n operators n and a complex number λ, +the conjugation does not change the order of the operators and does not conjugate complex +numbers: +(λ A1 · · · An |0⟩)t = λ ⟨0| (A1)t · · · (An)t. +(B.26) +However, it reverses radial ordering such that operators must be (anti-)commuted in radial +ordered expressions. +The BPZ product satisfies +⟨A, B⟩ = (−1)|A||B|⟨B, A⟩. +(B.27) +Moreover the inner product is non-degenerate, so +∀A : +⟨A|B⟩ = 0 +=⇒ +|B⟩ = 0. +(B.28) +Denoting by {|φr⟩} a complete basis of states, then the conjugate basis {⟨φc +r|} is defined +by the BPZ product as +⟨φc +r|φs⟩ = δrs. +(B.29) +We have +⟨φr|φc +s⟩ = (−1)|φr|δrs. +(B.30) +B.3.4 +Scalar field +The simplest matter CFT is a set of D scalar field Xµ(z, ¯z) such that the i∂Xµ and i¯∂Xµ +are of weight h = (1, 0) and h = (0, 1) +i∂Xµ = +� +n +αµ +n +zn+1 , +i¯∂Xµ = +� +n +¯αµ +n +¯zn+1 . +(B.31) +276 + +The commutation relations between the modes are: +[αµ +m, αν +n] = mδm+n,0ηµν, +[¯αµ +m, ¯αν +n] = mδm+n,0ηµν, +[αµ +m, ¯αν +n] = 0. +(B.32) +The zero-modes of both operators are equal and correspond to the (centre-of-mass) mo- +mentum +αµ +0 = ¯αµ +0 = +� +α′ +2 pµ. +(B.33) +The conjugate of pµ is the centre-of-mass position xµ: +[xµ, pν] = ηµν. +(B.34) +Vertex operators are defined by +Vk(z, ¯z) = :eik·X(z,¯z):, +h = ¯h = α′2k2 +4 +. +(B.35) +The scalar vacuum |k⟩ is annihilated by all positive-frequency oscillators and it is char- +acterized by its eigenvalue for the zero-mode operator +pµ |k⟩ = kµ |k⟩ , +∀n > 0 : +αµ +n |k⟩ = 0, +¯αµ +n |k⟩ = 0. +(B.36) +The vacuum is associated to the vertex operator Vk: +|k⟩ = Vk(0, 0) |0⟩ = eik·x |0⟩ . +(B.37) +The conjugate vacuum is +⟨k| pµ =⟨k| kµ, +⟨k| = |k⟩† , +⟨−k| = |k⟩t . +(B.38) +B.3.5 +Reparametrization ghosts +The reparametrization ghosts are described by an anti-commuting first-order system with +the parameters (Chapter 7 and table 7.1): +ϵ = 1, +λ = 2, +cgh = −26, +qgh = −3, +agh = −1. +(B.39) +We focus on the holomorphic sector. +The b and c ghosts have weights: +h(b) = 2, +h(c) = −1 +(B.40) +such that the mode expansions are: +b(z) = +� +n∈Z +bn +zn+2 , +c(z) = +� +n∈Z +cn +zn−1 , +(B.41a) +bn = +� +dz +2πi zn+1b(z), +cn = +� +dz +2πi zn−2c(z). +(B.41b) +The anti-commutators between the modes bn and cn read: +{bm, cn} = δm+n,0, +{bm, bn} = 0, +{cm, cn} = 0. +(B.42) +The energy–momentum tensor and the Virasoro modes are respectively: +T = −2 :b∂c: − :∂b c:, +(B.43a) +Lm = +� +n +� +n + m +� +:bm−ncn: = +� +n +(2m − n) :bncm−n:. +(B.43b) +277 + +The expression of the zero-mode is: +L0 = − +� +n +n :bnc−n: = +� +n +n :b−ncn:. +(B.44) +The commutators between the Ln and the ghost modes are: +[Lm, bn] = +� +m − n +� +bm+n, +[Lm, cn] = −(2m + n)cm+n. +(B.45) +In particular, L0 commutes with the zero-modes: +[L0, b0] = 0, +[L0, c0] = 0. +(B.46) +The anomalous global U(1) symmetry for the ghost number Ngh is generated by the +ghost current: +j = −:bc:, +Ngh,L = +� +dz +2πi j(z), +(B.47) +such that +Ngh(c) = 1, +Ngh(b) = −1. +(B.48) +Remember that Ngh = Ngh,L in the left sector, such that we omit the index L. The modes +of the ghost current are +jm = − +� +n +:bm−ncn: = − +� +n +:bncm−n:, +Ngh,L = j0 = − +� +n +:b−ncn:. +(B.49) +The commutator of the current modes with itself and with the Virasoro modes are: +[jm, jn] = m δm+n,0, +[Lm, jn] = −njm+n − 3 +2 m(m + 1)δm+n,0. +(B.50) +Finally, the commutators of the ghost number operator are: +[Ngh, b(w)] = −b(w), +[Ngh, c(w)] = c(w). +(B.51) +The level operators N b and N c and number operators N b +n and N c +n are defined as: +N b = +� +n>0 +n N b +n, +N c = +� +n>0 +n N c +n, +(B.52a) +N b +n = :b−ncn:, +N c +n = :c−nbn:. +(B.52b) +The commutator of the number operators with the modes are: +[N b +m, b−n] = b−nδm,n, +[N c +m, c−n] = c−nδm,n. +(B.53) +The OPE between the ghosts and different currents are: +c(z)b(w) ∼ +1 +z − w, +b(z)c(w) ∼ +1 +z − w, +b(z)b(w) ∼ 0, +c(z)c(w) ∼ 0, +(B.54a) +T(z)b(w) ∼ +2b(w) +(z − w)2 + ∂b(w) +z − w , +T(z)c(w) ∼ +−c(w) +(z − w)2 + ∂c(w) +z − w . +(B.54b) +j(z)b(w) ∼ − b(w) +z − w, +j(z)c(w) ∼ c(w) +z − w. +j(z)O(w) ∼ Ngh(O) O(w) +z − w, +(B.54c) +j(z)j(w) ∼ +1 +(z − w)2 . +(B.54d) +T(z)j(w) ∼ +−3 +(z − w)3 + +j(w) +(z − w)2 + ∂j(w) +z − w . +(B.54e) +278 + +any operator O(z) is defined by +The OPE (B.54e) implies that the ghost number is not conserved on a curved space: +N c − N b = 3 − 3g, +(B.55) +and leads to a shift between the ghost numbers on the plane and on the cylinder: +Ngh,L = N cyl +gh,L + 3 +2. +(B.56) +The SL(2, C) vacuum |0⟩ is defined by +∀n > −2 : +bn |0⟩ = 0, +∀n > 1 : +cn |0⟩ = 0. +(B.57) +The mode c1 does not annihilate the vacuum and the two degenerate energy vacua are: +| ↓⟩ := c1 |0⟩ , +| ↑⟩ := c0c1 |0⟩ . +(B.58) +The zero-point energy of these states is: +L0 | ↓⟩ = agh | ↓⟩ , +L0 | ↑⟩ = agh | ↑⟩ , +agh = −1. +(B.59) +The energy for the normal ordering of the different currents is: +Lm = +� +n +� +n − (1 − λ)m +� ⋆ +⋆bm−ncn +⋆ +⋆ + agh δm,0, +(B.60a) +jm = +� +n +⋆ +⋆bm−ncn +⋆ +⋆ + δm,0. +(B.60b) +The energy–momentum and ghost current zero-modes are explicitly: +L0 = +� +n +n +⋆ +⋆b−ncn +⋆ +⋆ + agh = �L0 − 1, +(B.61a) +Ngh,L = j0 = +� +n +⋆ +⋆b−ncn +⋆ +⋆ + 1 = � +Ngh,L + 1 +2 +� +N c +0 − N b +0 +� +− 3 +2, +(B.61b) +�L0 = N b + N c, +� +Ngh,L := +� +n>0 +� +N c +n − N b +n +� +. +(B.61c) +Then, one can straightforwardly compute the ghost number of the vacua: +Ngh |0⟩ = 0, +Ngh | ↓⟩ = | ↓⟩ , +Ngh | ↑⟩ = 2 | ↑⟩ . +(B.62) +Using (B.56) allows to write the ghost numbers on the cylinder: +N cyl +gh | ↓⟩ = −1 +2 | ↓⟩ , +N cyl +gh | ↑⟩ = 1 +2 | ↑⟩ . +(B.63) +The bn and cn are Hermitian: +b† +n = b−n, +c† +n = c−n. +(B.64) +The BPZ conjugates of the modes are: +bt +n = (±1)nb−n, +ct +n = −(±1)nc−n, +(B.65) +using I±(z) with (6.111). +279 + +The conjugates of the vacuum read: +| ↓⟩‡ =⟨0| c−1, +| ↑⟩‡ =⟨0| c−1c0. +(B.66) +The BPZ conjugates of the vacua are: +⟨↓ | := | ↓⟩t = ∓⟨0| c−1, +⟨↑ | := | ↑⟩t = ±⟨0| c0c−1. +(B.67) +We have the following relations: +⟨↓ | = ∓ | ↓⟩‡ , +⟨↑ | = ∓ | ↑⟩‡ . +(B.68) +The ghost are normalized with +⟨↑ | ↓⟩ =⟨↓ | c0 | ↓⟩ =⟨0| c−1c0c1 |0⟩ = 1, +(B.69) +which selects the minus sign in the BPZ conjugation. The conjugate of the ghost vacuum is +⟨0c| =⟨0| c−1c0c1. +(B.70) +Considering both the holomorphic and anti-holomorphic sectors, we introduce the com- +binations: +b± +n = bn ± ¯bn, +c± +n = 1 +2 (cn ± ¯cn). +(B.71) +The normalization of b± +m is chosen to match the one of L± +m (B.75), and the one of c± +m such +that +{b+ +m, c+ +n } = δm+n, +{b− +m, c− +n } = δm+n. +(B.72) +We have the following useful identities: +b− +n b+ +n = 2bn¯bn, +c− +n c+ +n = 1 +2 cn¯cn. +(B.73) +B.4 +Bosonic string +The BPZ conjugates of the scalar and ghost modes are +(αn)t = −(±1)n α−n, +(bn)t = (±1)n b−n, +(cn)t = −(±1)n c−n. +(B.74) +Combinations of holomorphic and anti-holomorphic modes: +L± +n = Ln ± ¯Ln, +b± +n = bn ± ¯bn, +c± +n = 1 +2 (cn ± ¯cn). +(B.75) +The closed string inner product is defined from the BPZ product by an additional inser- +tion of c− +0 +⟨A, B⟩ =⟨A| c− +0 |B⟩ , +(B.76) +while the open string inner product is equal to the BPZ product +⟨A, B⟩ = ⟨A|B⟩ . +(B.77) +The vacuum for the matter and ghosts is +|k, 0⟩ := |k⟩ ⊗ |0⟩ , +|k, ↓⟩ := |k⟩ ⊗ | ↓⟩ . +(B.78) +The vacuum is normalized as +open: +⟨k, ↓ | c0 |k, ↓⟩ =⟨k′, 0| c−1c0c1 |k, 0⟩ = (2π)Dδ(D)(k + k′), +(B.79a) +closed: +⟨k, ↓↓ | c0¯c0 |k, ↓↓⟩ =⟨k′, 0| c−1¯c−1c0¯c0c1¯c1 |k, 0⟩ = (2π)Dδ(D)(k + k′), +(B.79b) +280 + +Appendix C +Quantum field theory +In this appendix, we gather useful information on quantum field theories. The first section +describes how to compute with path integral with non-trivial measures, generalizing tech- +niques from finite-dimensional integrals. Then, we summarize the important concepts from +the BRST and BV formalisms. +C.1 +Path integrals +In this section, we explain how analysis, algebra and differential geometry are generalized +to infinite-dimensional vector spaces (fields). +C.1.1 +Integration measure +In order to construct a path integral for the field Φ, one needs to define a notion of distance +on the space of fields. The distance between a field Φ and a neighbouring field Φ + δΦ is +|δΦ|2 = G(Φ)(δΦ, δΦ), +(C.1) +where G is the (field-dependent) metric on the field tangent space (the field dependence will +be omitted when no confusion is possible). This induces a metric on the field space itself +|Φ|2 = G(Φ)(Φ, Φ), +(C.2) +from which the integration measure over the field space can be defined as +dΦ +� +det G(Φ). +(C.3) +Moreover, the field metric also defines an inner-product between two different elements of +the tangent space or field space: +(δΦ1, δΦ2) = G(Φ)(δΦ1, δΦ2), +(Φ1, Φ2) = G(Φ)(Φ1, Φ2). +(C.4) +Remark C.1 (Metric in component form) If one has a set of spacetime fields Φa(x), +then a local norm is defined by +|δΦa|2 = +� +dx ρ(x)γab +� +Φ(x) +� +δΦa(x)δΦb(x), +(C.5) +which means that the metric in component form is +Gab(x, y)(Φ) = δ(x − y)ρ(x)γab +� +Φ(x) +� +. +(C.6) +281 + +Locality means that all fields are evaluated at the same point. +On a curved space, it is +natural to write γ only in terms of the metric g and to set ρ(x) = +� +det g(x), such that the +inner-product is diffeomorphism invariant. +Since a Gaussian integral is proportional to the squareroot of the operator determinant, +the integration measure can be determined by considering the Gaussian integral over the +tangent space: +� +dδΦ e−G(Φ)(δΦ,δΦ) = +1 +� +det G(Φ) +. +(C.7) +Note that one needs to work on the tangent space because G(Φ) can depend on the field, +which means that the integral +� +dΦ e−G(Φ)(Φ,Φ). +(C.8) +is not Gaussian. +Having constructed the Gaussian measure with respect to the metric G(Φ), it is now +possible to consider the path integral of general functional F of the fields: +� +dΦ +� +det G(Φ) F(Φ). +(C.9) +The (effective) action S(Φ) provides a natural metric on the field space by defining +√ +det G = +e−S, or +S = −1 +2 tr ln G(Φ). +(C.10) +However, it can be simpler to work with a Gaussian measure by considering only the quad- +ratic terms in S, and expanding the rest in a power series. +In particular, the partition +function is defined from the classical action Scl by +Z = +� +dΦ e−Scl(Φ). +(C.11) +Given an operator D, its adjoint D† is defined with respect to the metric as +G(δΦ, DδΦ) = G(D†δΦ, δΦ). +(C.12) +The free-field measure is such that the metric on the field space is independent from the +field itself: G(X) = G0. In particular, this implies that the metric is flat and its determinant +can be absorbed in the measure, setting det G0 = 1. In this case, the measure is invariant +under shift of the field: +Φ → Φ + ε +(C.13) +such that +� +dΦ e− 1 +2 |Φ+ε|2 = +� +dΦ e− 1 +2 |Φ|2. +(C.14) +This property allows to complete squares and shift integration variables (for example to +generate a perturbative expansion and to derive the propagator). +Computation – Equation (C.14) +� +dΦ e− 1 +2 |Φ+ε|2 = +� +d�Φ det δΦ +δ�Φ +e− 1 +2 |� +Φ| +2 += +� +d�Φ e− 1 +2 |� +Φ| +2 +(C.15) +The first equality follows by setting �Φ = Φ + ε, and the result (C.14) follows by the +redefinition �Φ = Φ. +282 + +C.1.2 +Field redefinitions +Under a field redefinition Φ → Φ′, the norm and the measure are invariant: +dΦ +� +det G(Φ) = d�Φ +� +det �G(�Φ), +G(Φ)(δΦ, δΦ) = �G(�Φ)(δ�Φ, δ�Φ). +(C.16) +Conversely, one can find the Jacobian J(Φ, �Φ) between two coordinate systems by writing +dΦ = J(Φ, �Φ)d�Φ, +J(Φ, �Φ) = +����det ∂Φ +∂�Φ +���� = +� +det �G(�Φ) +det G(Φ). +(C.17) +If the measure of the initial field coordinate is normalized such that det G = 1, or equivalently +� +dδΦ e−|δΦ|2 = 1, +(C.18) +one can determine the Jacobian by performing explicitly the integral +J(�Φ)−1 = +� +dδ�Φ e−� +G(δ� +Φ,δ� +Φ). +(C.19) +Remark C.2 (Identity of the Jacobian for Φ and δΦ) The Jacobian agrees on the space +of fields and on its tangent space. This is most simply seen by using a finite-dimensional +notation: considering the coordinates xµ and a vector v = vµ∂µ, the Jacobian for changing +the coordinates to ˜xµ is equivalently +J = det ∂˜xµ +∂xµ = det ∂˜vµ +∂vµ +(C.20) +since the vector transforms as +˜vµ = vν ∂˜xµ +∂xν . +(C.21) +C.1.3 +Zero-modes +A zero-mode Φ0 of an operator D is a field such that +DΦ0 = 0. +(C.22) +In the definition of the path integral over the space of fields Φ, the measure is defined +over the complete space. However, this will lead respectively to a divergent or vanishing +integral if the field is bosonic or fermionic, because the integration over the zero-modes can +be factorized from the rest of the integral. Writing the field as +Φ = Φ0 + Φ′, +(Φ0, Φ′) = 0, +(C.23) +where Φ′ is orthogonal to the zero-mode Φ0, a Gaussian integral of an operator D reads: +Z[D] = +� +dΦ +√ +det G e− 1 +2 (Φ,DΦ) = +�� +dΦ0 +� � +dΦ′ e− 1 +2 (Φ′,DΦ′) +(C.24) +A first solution could be to simply strip the first factor (for example, by absorbing it in the +normalization), but this is not satisfactory. In particular, the partition function with source +Z[D, J] = +� +dΦ +√ +det G e− 1 +2 (Φ,DΦ)−(J,Φ) +(C.25) +283 + +will depend on the zero-modes through the sources. +But, since the zero-modes are still +singled out, it is interesting to factorize the integration +Z[D, J] = +� +dΦ0 e−(J,Φ0) +� +dΦ′ e− 1 +2 (Φ′,DΦ′)−(J,Φ′) +(C.26) +and to understand what makes it finite. Ensuring that zero-modes are correctly inserted +is an important consistency and leads to powerful arguments. Especially, this can help to +guess an expression when it cannot be derived easily from first principles. +To exemplify the problem, consider the cases where there is a single constant zero-mode +denoted as x (bosonic) or θ (fermionic). The integral over x is infinite: +� +dx = ∞. +(C.27) +Oppositely, the integral of a Grassmann variable θ vanishes: +� +dθ = 0. +(C.28) +A Grassmann integral satisfies also +� +dθ θ = +� +dθ δ(θ) = 1, +(C.29) +such that an integral over a zero-mode does not vanish if there one zero-mode in the integrand +(due to the Grassmann nature of θ, the integrand can be at most linear). By analogy with +the fermionic case, a possibility for getting a finite bosonic integral is to insert a delta +function: +� +dx δ(x) = 1. +(C.30) +We will see that this is exactly what happens for the ghosts and super-ghosts in (super)string +theories. +Since ker D is generally finite-dimensional, it is interesting to decompose the zero-mode +on a basis and to integrate over the coefficients in order to obtain a finite-dimensional +integral. Writing the zero-mode as +θ0(x) = θ0iψi(x), +ker D = Span{ψi} +(C.31) +where the coefficients θ0i are constant Grassmann numbers, the change of variables θ → +(θ0i, θ′) implies: +dθ = +1 +� +det(ψi, ψj) +dθ′ +n +� +i=1 +dθ0i, +(C.32) +where n = dim ker D. +Next, according to the discussion above, one can ask if it is possible to rewrite an integ- +ration over dθ′ in terms of an integration over dθ together with zero-mode insertions. This +is indeed possible and one finds: +dθ +n +� +i=1 +θ(xi) = +det ψi(xj) +� +det(ψi, ψj) +dθ′. +(C.33) +284 + +Computation – Equation (C.32) +1 = +� +dθ e−|θ|2 = +� +dθ′dθ0 e−|θ|2−|θ0|2 += J +� +dθ′ � +i +dθ0i e−|θ′|2−|θ0iψi|2 = J +� +det(ψi, ψj) +Computation – Equation (C.33) +The simplest approach is to start with the LHS. This formula is motivated from the +previous discussion: if the integration measure contains n zero-modes, it will vanish +unless there are n zero-mode insertions. Moreover, one can replace each of them by the +complete field since only the zero-mode part can contribute: +� +dθ0 +n +� +j=1 +θ(xj) = +� +dθ0 +n +� +j=1 +θ0(xj) = +1 +� +det(ψi, ψj) +� +dnθ0i +n +� +j=1 +� +θ0iψi(xj) +� += +det ψi(xj) +� +det(ψi, ψj) +� � +i +dθ0i θ0i = +det ψi(xj) +� +det(ψi, ψj) +. +The third equality follows by developing the product and ordering the θ0i: minus signs +result from anticommuting the θ0i such that one gets the determinant of the basis +elements. +C.2 +BRST quantization +Consider an action Sm[φi] which depends on some fields φi subject to a gauge symmetry: +δφi = ϵaδaφi = ϵaRi +a(φ), +(C.34) +where ϵa are the (local) bosonic parameters, such that the action is invariant +ϵaδaSm = 0. +(C.35) +The gauge transformations form a Lie algebra with structure coefficients f c +ab +[δa, δb] = f c +abδc. +(C.36) +It is important 1) that the algebra closes off-shell (without using the equations of motion), +2) that the structure coefficients are field independent and 3) that the gauge symmetry is +irreducible (each gauge parameter is independent). +Remark C.3 (Interpretation of the Ri +a matrices) If the φi transforms in a represent- +ation R of the gauge group, then the transformation is linear in the field +Ri +a(φ) = (T R +a )i +jφj, +(C.37) +with T R +a the generators in the representation R. But, in full generality, this is not the case: +for example the gauge fields Aa +µ do not transform in the adjoint representation even if they +carry an adjoint index (only the field strength does), and in this case +Rb +aµ = δb +a∂µ + f c +abAb +µ. +(C.38) +When the fields φi form a non-linear sigma models, the Ri +a(φ) correspond to Killing +vectors of the target manifold. +285 + +In order to fix the gauge symmetry in the path integral +Z = Ω−1 +gauge +� +dφi e−Sm, +(C.39) +gauge fixing conditions must be imposed: +F A(φi) = 0. +(C.40) +Indeed, without gauge fixing, the integration is performed over multiple identical configura- +tions and the result diverges. The index A is different from the gauge index a because they +can refer to different representations, but for the gauge fixing to be possible they should run +over as many values. +Next, ghost fields ca (fermionic) are introduced for every gauge parameter, anti-ghosts +bA (fermionic) and auxiliary (Nakanishi–Laudrup) fields BA (bosonic) for every gauge con- +dition. The gauge-fixing and ghost actions are then defined by +Sgh = bAca δaF A(φi), +(C.41a) +Sgf = −i BAF A(φi) +(C.41b) +such that the original partition function is equivalent to +Z = +� +dφi dbA dca dBA e−Stot +(C.42) +where +Stot = Sm + Sgf + Sgh. +(C.43) +The total action is invariant +δϵStot = 0. +(C.44) +under the (global) BRST transformations +δϵφi = iϵ caδaφi, +δϵca = − i +2 ϵ f a +bccbcc, +δϵbA = ϵ BA, +δϵBA = 0, +(C.45) +where ϵ is an anti-commuting constant parameter. +Note that the original action Sm is +invariant by itself since the transformation acts like a gauge transformation with parameter +ϵca. The transformation of ca follows because it transforms in the adjoint representation of +the gauge group. Direct computations show that this transformation is nilpotent +δϵδϵ′ = 0. +(C.46) +These transformations are generated by a (fermionic) charge QB called the BRST charge +δϵφi = i [ϵQB, φi] +(C.47) +and similarly for the other fields (stripping the ϵ outside the commutator turns it to an +anticommutator if the field is fermionic). Taking the ghosts to be Hermitian leads to an +Hermitian charge. +An important consequence is that the two additional terms of the action can be rewritten +as a BRST exact terms +Sgf + Sgh = {QB, bAF A}. +(C.48) +A small change in the gauge-fixing condition δF leads to a variation of the action +δS = {QB, bAδF A}. +(C.49) +286 + +The BRST charge should commute with the Hamiltonian in order to be conserved: this +should hold in particular when changing the gauge fixing condition +[QB, {QB, bAδF A}] = 0 +=⇒ +Q2 +B = 0. +(C.50) +Some vocabulary is needed before proceeding further. A state |ψ⟩ is said to be BRST +closed if it is annihilated by the BRST charge +|ψ⟩ closed +⇐⇒ +|ψ⟩ ∈ ker QB +⇐⇒ +QB |ψ⟩ = 0. +(C.51) +States which are in the image of QB (i.e. they can be written as QB applied on some other +states) are said to be exact +|ψ⟩ exact +⇐⇒ +|ψ⟩ ∈ Im QB +⇐⇒ +∃ |χ⟩ : |ψ⟩ = QB |χ⟩ . +(C.52) +The cohomology H(QB) of QB is the set of closed states which are not exact +|ψ⟩ ∈ H(QB) +⇐⇒ +|ψ⟩ ∈ ker QB, +∄ |χ⟩ : |ψ⟩ = QB |χ⟩ . +(C.53) +Hence the cohomology corresponds to +H(QB) = ker QB +Im QB +. +(C.54) +Two elements of the cohomology differing by an exact state are in the same equivalence class +|ψ⟩ ≃ |ψ⟩ + QB |χ⟩ . +(C.55) +Considering the S-matrix ⟨ψf|ψi⟩ between a set of physical initial states ψi and final +states ψf, a small change in the gauge-fixing condition leads to +δF ⟨ψf|ψi⟩ =⟨ψf| {QB, bAδF A} |ψi⟩ +(C.56) +after expanding the exponential to first order. Since the S-matrix should not depend on the +gauge this implies that a physical state ψ must be BRST closed (i.e. invariant) +QB |ψ⟩ = 0. +(C.57) +Conversely, this implies that any state of the form QB |χ⟩ cannot be physical because it is +orthogonal to every physical state |ψ⟩ +⟨ψ| QB |χ⟩ = 0. +(C.58) +This implies in particular that the amplitudes involving |ψ⟩ and |ψ⟩ + QB |χ⟩ are identical, +and any amplitude for which an external state is exact vanishes. As a conclusion, physical +states are in the BRST cohomology +|ψ⟩ physical +⇐⇒ +|ψ⟩ ∈ H(QB). +(C.59) +If there is a gauge where the ghosts decouple from the matter field, then the invariance +of the action and of the S-matrix under changes of the gauge fixing ensures that this state- +ment holds in any gauge (but, one still need to check that the gauge preserves the other +symmetries). If such a gauge does not exist, then one needs to employ other methods to +show the desired result. +Note that BA can be integrated out by using its equations of motion +δF A +δφi BA = −δSm +δφi , +(C.60) +287 + +and this modifies the BRST transformation of the anti-ghost to +δϵbA = −ϵ +�δF A +δφi +�−1 δSm +δφi . +(C.61) +It is also possible to introduce a term +{QB, bABBM AB} = i BAM ABBB +(C.62) +for any constant matrix M AB. Since this is also a BRST exact term, the amplitudes are +not affected. Integrating over BA produces a Gaussian average instead of a delta function +to fix the gauge. +In the previous discussion, the BRST symmetry was assumed to originate from the +Faddeev–Popov gauge fixing. But, in fact, it is possible to start directly with an action of +the form +S[φ, b, c, B] = S0[φ] + QBΨ[φ, b, c, B] +(C.63) +where Ψ has ghost number −1. It can be proven that this is the most general action invariant +under the BRST transformations (C.45). This can describe gauge fixed action which cannot +be described by the Faddeev–Popov procedure: in particular, the latter yields actions which +are quadratic in the ghost fields (by definition of the Gaussian integral representation of the +determinant), but this does not exhaust all the possibilities. For example, the background +field method applied to Yang–Mills theory requires using an action quartic in the ghosts. +In this section, several hypothesis have been implicit (off-shell closure, irreducibility and +constant structure coefficients). If one of them breaks, then it is necessary to employ the +more general BV formalism. +C.3 +BV formalism +The Batalin–Vilkovisky (BV, or also field–antifield) formalism is the most general framework +to quantize theories with a gauge symmetry. While the BRST formalism (Appendix C.2) +is sufficient to describe simple systems, it breaks down when the structure of the gauge +symmetry is more complicated, for example in systems implying gravity. The BV formalism +is required in the three following cases (which can occur simultaneously): +1. the gauge algebra is open (on-shell closure); +2. the structure coefficients depend on the fields; +3. the gauge symmetry is reducible (not all transformations are independent). +The BV formalism is also useful for standard gauge symmetries to demonstrate renormaliz- +ability and to deal with anomalies. +As explained in the previous section, the ghosts and the BRST symmetry are crucial to +ensure the consistency of the gauge theory. The idea of the BV formalism is to put on an +equal footing the physical fields and all the required auxiliary and ghost fields (before gauge +fixing). The introduction of antifields – one for each of the fields – and the description of the +full quantum dynamics in terms of a quantum action (constrained by the quantum master +equation) ensure the consistency of the system. Additional benefits are the presence of a +(generalized) BRST symmetry, the existence of a Poisson structure (which allows to bring +concepts from the Hamiltonian formalism), the covariance of the formalism and the simple +interpretation of counter-terms as corrections to the classical action. +For giving a short intuition, the BV formalism can be interpreted as providing a (anti)ca- +nonical structure in the Lagrangian formalism, the role of the Hamiltonian being played by +the action. +288 + +C.3.1 +Properties of gauge algebra +Before explaining the BV formalism, we review the situations listed above. The classical +action for the physical fields φi is denoted by S0[φ] and the associated equations of motion +by +Fi(φ) = ∂S0 +∂φi . +(C.64) +Then, a gauge algebra is open and has field-dependent structure coefficients F c +ab(φ) if: +[Ta, Tb] = F c +ab(φ)Tc + λi +abFi(φ). +(C.65) +On-shell, Fi = 0 and the second term is absent, such that the algebra closes. The fields +themselves are constants from the point of view of the gauge algebra, but their presences in +the structure coefficients complicate the analysis of the theory. Moreover, the path integral +is off-shell and for this reason one needs to take into account the last term. +Finally, the gauge algebra can be reducible: in brief, it means that there are gauge +invariances associated to gauge parameters – and correspondingly ghosts for ghosts –, and +this recursively. Since there is one independent ghost for each generator, there are too many +ghosts if the generators are not all independent, and there is a remnant gauge symmetry for +the ghost fields (in the standard Faddeev–Popov formalism, the ghosts are not subject to +any gauge invariance). This originates from relations between the generators Ri +a: denoting +by m0 the number of level-0 gauge transformations, the number of independent generators +is rank Ri +a. Then, the +m1 = m0 − rank Ri +a +(C.66) +relations between the generators translate into a level-1 gauge invariance of the ghosts. This +symmetry can be gauge fixed by performing a second time the Faddeev–Popov procedure, +yielding commuting ghosts. This symmetry can also be reducible, and the procedure can +continue without end. If one finds that the gauge invariance at level n = ℓ is irreducible, one +says that the gauge invariance is ℓ-reducible. If this does not happen, one defines ℓ = ∞. +The number of generators at level n is denoted by mn. +Example C.1 – p-form gauge theory +A p-form gauge theory is written in terms of a gauge field Ap with a a gauge invariance +δAp = dλp−1. +(C.67) +But, due to the nilpotency of the derivative, deformations of the gauge parameter +satisfying +δλp−1 = dλp−2 +(C.68) +does not translate into a gauge invariance of Ap. Similarly from this should be excluded +the transformation +δλp−2 = dλp−3, +(C.69) +and so on until one reaches the case p = 0. Hence, a p-form field has a p-reducible +gauge invariance. +C.3.2 +Classical BV +Denoting the fields collectively as +ψr = {φi, BA, bA, ca}, +(C.70) +289 + +the simplest BV action reads +S[ψr, ψ∗ +r] = S0[φ] + QBψr ψ∗ +r +(C.71) +with the antifields +ψ∗ +r = {φ∗ +i , BA∗, bA∗, c∗ +a}. +(C.72) +The action (C.63) is recovered by writing +ψ∗ +r = ∂Ψ +∂ψr . +(C.73) +This indicates that the general BRST formalism could be rephrased in the BV language. +But, in the same way that the BRST formalism generalizes the Faddeev–Popov formalism, +it is in turn generalized by the BV formalism. Indeed, the above action is linear in the +antifields: this constraint is not required and one can write more general actions. In the rest +of this section, we explain how this works at the level of the action (classical level) and how +the sets of fields and antifields are defined. +Consider a set of physical fields φi with the gauge invariance +δφi = ϵa0 +0 Ri +a0(φi). +(C.74) +Then, associate a ghost field ca0 to each of the gauge parameters ϵa0. If the gauge symmetry +is reducible, a new gauge invariance is associated to the ghosts +δca0 +0 = ϵa1 +1 Ra0 +a1(φi, ca0). +(C.75) +This structure is recurring and the ghosts of the level-n gauge invariance are denoted by can +and they satisfy +δcan +n = ϵan+1 +n+1 Ran +an+1(φi, ca0 +0 , . . . , can +n ). +(C.76) +Thus, the set of fields is +ψr = {can +n }n=−1,...,ℓ, +c−1 := φ. +(C.77) +A ghost number is introduced +Ngh(φi) = 0, +Ngh(can +n ) = n + 1, +(C.78) +and the Grassmann parity of the ghosts is defined to be opposite (resp. identical) of the +parity of the associated gauge parameter for even (resp. odd) n +|cn| = |ϵan +n | + n + 1. +(C.79) +To each of these fields is associated an antifield ψ∗ +r of opposite parity as ψr and such that +their ghost numbers sum to −1 +Ngh(ψ∗ +r) = −1 − Ngh(ψr), +|ψ∗ +r| = −|ψr|. +(C.80) +The fields and antifields together are taken to define a graded symplectic structure +ω = +� +r +dψr ∧ dψ∗ +r +(C.81) +with respect to which they are conjugated to each other +(ψr, ψ∗ +s) = δrs, +(ψr, ψs) = 0, +(ψ∗ +r, ψ∗ +s) = 0. +(C.82) +290 + +The antibracket (graded Poisson bracket) (·, ·) reads +(A, B) = ∂RA +∂ψr +∂LB +∂ψ∗r +− ∂RA +∂ψ∗r +∂LB +∂ψr , +(C.83) +where the L and R indices indicate left and right derivatives. It is graded symmetric, which +means +(A, B) = −(−1)(|A|+1)(|B|+1)(B, A). +(C.84) +It also satisfies a graded Jacobi identity and the property +Ngh((A, B)) = Ngh(A) + Ngh(B) + 1, +|(A, B)| = |A| + |B| + 1 +mod 2. +(C.85) +Moreover, the antibracket acts as a derivative +(A, BC) = (A, B)C + (−1)|B|C(A, C)B. +(C.86) +The dynamics of the theory is described by the (classical) master action S[ψr, ψ∗ +r] which +satisfies +Ngh(S) = 0, +|S| = 0. +(C.87) +In order to reproduce correctly the dynamics of the classical system without ghosts, this +action is required to satisfy the boundary condition +S[ψr, ψ∗ +r = 0] = S0[φi], +∂L∂RS +∂c∗ +n−1,an−1∂can +n +���� +ψ∗=0 += Ran−1 +an +. +(C.88) +Indeed, if the antifields are set to zero, the ghost fields cannot appear because they all have +positive ghost numbers and it is not possible to build terms with vanishing ghost numbers +from them. +In analogy with the Hamiltonian formalism, the master action can be used as the gen- +erator of a global fermionic symmetry, and inspection will show that it corresponds to a +generalization of the BRST symmetry. Writing the generalized and classical BRST operator +as s, the transformations of the fields and antifields read +δθψr = θ sψr = −θ (S, ψr) = θ ∂RS +∂ψ∗r +, +(C.89a) +δθψ∗ +r = θ sψ∗ +r = −θ (S, ψ∗ +r) = −θ ∂RS +∂ψr , +(C.89b) +where θ is a constant Grassmann parameter. The variation of a generic functional F[ψr, ψ∗ +r] +is +δθF = θ sF = −θ (S, F). +(C.90) +For the BRST transformation to be a symmetry of the action, the action must satisfy the +classical master equation +(S, S) = 0. +(C.91) +This equation can easily be solved by expanding S in the ghosts: the various terms can be +interpreted in terms of properties of the gauge algebra. Then, the Jacobi identity used with +two S and an arbitrary functional gives +(S, (S, F)) = 0 +(C.92) +and this implies that the transformation is nilpotent +s2 = 0. +(C.93) +291 + +A classical observable O satisfies +sO = 0. +(C.94) +Due to the BRST symmetry, the action is not uniquely defined and the action +S′ = S + (S, δF) +(C.95) +also satisfies the master equation, where δF is arbitrary up to the condition Ngh(δF) = −1. +This can be interpreted as the action S in a new coordinate system (ψ′r, ψ′∗ +r ) with +ψ′ = ψ − δF +δψ∗ , +ψ′∗ = ψ∗ + δF +δψ +(C.96) +such that +S′[ψ, ψ∗] = S +� +ψ − δF +δψ∗ , ψ∗ + δF +δψ +� +. +(C.97) +Indeed, for F = F[ψ, ψ∗], one has +S′[ψ, ψ∗] = S[ψ, ψ∗] + (S, ψ)δF +δψ + (S, ψ∗) δF +δψ∗ = S[ψ, ψ∗] − ∂RS +∂ψ∗ +δF +δψ + ∂RS +∂ψ +δF +δψ∗ . +(C.98) +It can be shown that this transformation preserves the antibracket and the master equation +(ψ′r, ψ′∗ +s ) = δrs, +(S′, S′) = 0. +(C.99) +More generally, any transformation preserving the antibracket is called an (anti)canonical +transformation. One can also consider generating functions depending on both the old and +new coordinates, as is standard in the Hamiltonian formalism. Under a transformation, any +object depending on the coordinates changes as +G′ = G + (δF, G). +(C.100) +One can consider finite transformation without problems. +In order to perform the gauge fixing, one needs to eliminate the antifields. A convenient +condition is +SΨ[ψr] = S +� +ψr, ∂Ψ +∂ψr +� +, +ψ∗ +r = ∂Ψ +∂ψr , +(C.101) +where Ψ[ψr] is called the gauge fixing fermion and satisfies +Ngh(Ψ) = −1, +|Ψ| = 1. +(C.102) +From the discussion on coordinate transformations this amounts to work in new coordinates +where ψ′∗ +r = 0. But such a function Ψ cannot be built from the fields because they all have +positive ghost numbers. One needs to introduce trivial pairs of fields. +A trivial pair (B, ¯c) is defined by the properties +|B| = −|¯c|, +Ngh(B) = Ngh(¯c) + 1, +(C.103a) +s¯c = B, +sB = 0 +(C.103b) +and the new action reads +¯S = S[ψr, ψ∗ +r] − B¯c∗ +(C.104) +(the position dependence is kept implicit). In this context ψr and ψ∗ +r are sometimes called +minimal variables. From this, one learns that +( ¯S, ¯S) = (S, S) = 0. +(C.105) +292 + +At level-0, one introduces the pair +(B0a0, ¯c0a0) := (B0 +0a0, ¯c0 +0a0) +(C.106) +and the associated antifields. The field ¯c0 := b is the Faddeev–Popov anti-ghost associated +to c0 and the trivial pair satisfies +|B0| = |ϵ0|, +|¯c0| = −|ϵ0|, +Ngh(B0) = 0, +Ngh(c0) = −1. +(C.107) +For the level 1, two additional pairs are introduced: +(B0 +1a1, ¯c0 +1a1), +( ¯B1a1 +1 +, c1a1 +1 +) +(C.108) +and the corresponding antifields. The motivation for adding an additional pair is that the +level-0 pair only fixes m0 − m1 of the generators: the additional m1 extra-ghosts c1a1 +1 +can +be fixed by the residual level-0 symmetry. The first level-1 pair fixes the level-1 symmetry. +Then, the gauge fixed action enjoys a BRST symmetry acting only on the fields +δθψr = θ sψr = θ ∂RS +∂ψ∗r +���� +ψ∗ +r =∂rΨ +. +(C.109) +Note that this BRST operator is generically nilpotent only on-shell +s2 ∝ eom. +(C.110) +C.3.3 +Quantum BV +At the quantum level, one considers the path integral +Z = +� +dψrdψ∗ +r e−W [ψr,ψ∗ +r ]/ℏ +(C.111) +where W is called the quantum master action. The reason for distinguishing it from the +classical master action S is that the measure is not necessarily invariant by itself under +the generalized BRST transformation – this translates into a non-gauge invariance of the +measure of the physical fields, i.e. a gauge anomaly. +Quantum BRST transformation are generated by the quantum BRST operator σ +δθF = θ σF = (W, F) − ℏ ∆F, +(C.112) +where +∆ = ∂R +∂ψ∗r +∂L +∂ψr . +(C.113) +Then, the path integral is invariant if W satisfies the quantum master equation +(W, W) − 2ℏ∆W = 0, +(C.114) +which can also be written as +∆e−W/ℏ = 0. +(C.115) +This can be interpreted as the invariance of Z under changes of coordinates: indeed one +finds that +δW = 1 +2(W, W), +(C.116) +and the integration measure picks a Jacobian +sdet J ∼ 1 + ∆W. +(C.117) +293 + +In the limit ℏ → 0, one recovers the classical master equation. More generally, the action +can be expanded in powers of ℏ +W = S + +� +p≥1 +ℏpWp. +(C.118) +Observables are given by operators O[ψ, ψ∗] invariant under σ: +σO = 0, +(C.119) +which ensures that the expectation value is invariant under changes of Ψ +δ⟨O⟩ = 0. +(C.120) +Note that if O depends just on ψ the condition reduces to sO = 0, but generically there is +no such operators (except constants) satisfying this condition for open algebra. +Consider the gauge fixed integral +Z = +� +dψr e−WΨ[ψr], +WΨ[ψr] = W +� +ψr, ∂Ψ +∂ψr +� +. +(C.121) +Varying the gauge fixing fermion by δΨ gives +Z = +� +dψr e−WΨ[ψr] +�∂RS +∂ψ∗r +� +ψ∗=∂ψΨ +∂(δΨ) +∂ψr . +(C.122) +Integrating by part gives the quantum master equation. +C.4 +Suggested readings +• Manipulations of functional integral are given in [100, sec. 15.1, 22.1, 172, chap. 14, +191, 53]. +• Zero-modes are discussed in [23]. +• A general summary of path integrals for bosonic and fermionic fields can be found +in [193, app. A]. +• BRST formalism: most QFT books contain an introduction, more complete references +are [251, chap. 15, 247, 105]; +• BV formalism [251, chap. 15, 88, 93, 247, 105, 49] (several explicit examples are given +in [93, sec. 3], see [12, 231, 254] for more specific details). +294 + +Bibliography +[1] +M. Asano and M. Kato. ‘New Covariant Gauges in String Field Theory’. Progress of +Theoretical Physics 117.3 (Mar. 2007), pp. 569–587. doi: 10.1143/PTP.117.569. +arXiv: hep-th/0611189. +[2] +M. Asano and M. Kato. ‘General Linear Gauges and Amplitudes in Open String +Field Theory’. Nuclear Physics B 807.1-2 (Jan. 2009), pp. 348–372. doi: 10.1016/ +j.nuclphysb.2008.09.004. arXiv: 0807.5010. +[3] +M. Asano and M. Natsuume. ‘The No-Ghost Theorem for String Theory in Curved +Backgrounds with a Flat Timelike Direction’. Nuclear Physics B 588.1-2 (Nov. 2000), +pp. 453–470. doi: 10.1016/S0550-3213(00)00495-8. arXiv: hep-th/0005002. +[4] +Y. Baba, N. Ishibashi and K. Murakami. ‘Light-Cone Gauge String Field Theory +in Noncritical Dimensions’. Journal of High Energy Physics 2009.12 (Dec. 2009), +pp. 010–010. doi: 10.1088/1126-6708/2009/12/010. arXiv: 0909.4675. +[5] +I. Bars and C. N. Pope. ‘Anomalies in Super P-Branes’. Classical and Quantum +Gravity 5.9 (1988), p. 1157. doi: 10.1088/0264-9381/5/9/002. +[6] +I. Bars, C. N. Pope and E. Sezgin. ‘Central Extensions of Area Preserving Membrane +Algebras’. Physics Letters B 210.1 (Aug. 1988), pp. 85–91. doi: 10.1016/0370- +2693(88)90354-1. +[7] +I. Bars. ‘First Massive Level and Anomalies in the Supermembrane’. Nuclear Physics +B 308.2 (Oct. 1988), pp. 462–476. doi: 10.1016/0550-3213(88)90573-1. +[8] +I. Bars. ‘Is There a Unique Consistent Unified Theory Based on Extended Objects?’ +(Aug. 1988), pp. 0693–698. url: https://inspirehep.net/literature/264587. +[9] +I. Bars. ‘Issues of Topology and the Spectrum of Membranes’ (Jan. 1990), pp. 0209– +212. url: https://inspirehep.net/literature/285041. +[10] +I. Bars. ‘Membrane Symmetries and Anomalies’. Nuclear Physics B 343.2 (Oct. 1990), +pp. 398–417. doi: 10.1016/0550-3213(90)90476-T. +[11] +I. Bars, C. N. Pope and E. Sezgin. ‘Massless Spectrum and Critical Dimension of the +Supermembrane’. Physics Letters B 198.4 (Dec. 1987), pp. 455–460. doi: 10.1016/ +0370-2693(87)90899-9. +[12] +I. A. Batalin and G. A. Vilkovisky. ‘Quantization of Gauge Theories with Linearly De- +pendent Generators’. Physical Review D 28.10 (Nov. 1983), pp. 2567–2582. doi: 10. +1103/PhysRevD.28.2567. +[13] +T. Bautista, A. Dabholkar and H. Erbin. ‘Quantum Gravity from Timelike Liouville +Theory’. Journal of High Energy Physics 2019.10 (Oct. 2019), p. 284. doi: 10.1007/ +JHEP10(2019)284. arXiv: 1905.12689. +[14] +K. Becker, M. Becker and J. H. Schwarz. String Theory and M-Theory: A Modern +Introduction. 1st edition. Cambridge University Press, Dec. 2006. +295 + +[15] +A. A. Belavin and V. G. Knizhnik. ‘Algebraic Geometry and the Geometry of Quantum +Strings’. Physics Letters B 168.3 (Mar. 1986), pp. 201–206. doi: 10.1016/0370- +2693(86)90963-9. +[16] +A. Belopolsky and B. Zwiebach. ‘Who Changes the String Coupling ?’ Nuclear Phys- +ics B 472.1-2 (July 1996), pp. 109–138. doi: 10.1016/0550- 3213(96)00203- 9. +arXiv: hep-th/9511077. +[17] +O. Bergman and B. Zwiebach. ‘The Dilaton Theorem and Closed String Backgrounds’. +Nuclear Physics B 441.1-2 (May 1995), pp. 76–118. doi: 10.1016/0550-3213(95) +00022-K. arXiv: hep-th/9411047. +[18] +N. Berkovits. ‘Super-Poincare Invariant Superstring Field Theory’. Nuclear Physics B +450.1-2 (Sept. 1995), pp. 90–102. doi: 10.1016/0550-3213(95)00259-U. arXiv: hep- +th/9503099. +[19] +N. Berkovits, Y. Okawa and B. Zwiebach. ‘WZW-like Action for Heterotic String +Field Theory’. Journal of High Energy Physics 2004.11 (Nov. 2004), pp. 038–038. +doi: 10.1088/1126-6708/2004/11/038. arXiv: hep-th/0409018. +[20] +A. Bilal. ‘Remarks on the BRST-cohomology for cM > 1 matter coupled to "Liouville +gravity"’. Physics Letters B 282.3-4 (May 1992), pp. 309–313. doi: 10.1016/0370- +2693(92)90644-J. arXiv: hep-th/9202035. +[21] +A. Bilal and C. de Lacroix. ‘2D Gravitational Mabuchi Action on Riemann Sur- +faces with Boundaries’. Journal of High Energy Physics 2017.11 (Nov. 2017), p. 154. +doi: 10.1007/JHEP11(2017)154. arXiv: 1703.10541. +[22] +A. Bilal and L. Leduc. ‘2D Quantum Gravity on Compact Riemann Surfaces with +Non-Conformal Matter’. Journal of High Energy Physics 2017.01 (2017), p. 089. +doi: 10.1007/JHEP01(2017)089. arXiv: 1606.01901. +[23] +S. K. Blau, M. Visser and A. Wipf. ‘Determinants, Dirac Operators, and One-Loop +Physics’. International Journal of Modern Physics A 04.06 (Apr. 1989), pp. 1467– +1484. doi: 10.1142/S0217751X89000625. +[24] +R. Blumenhagen, D. Lüst and S. Theisen. Basic Concepts of String Theory. 2013 +edition. Springer, Nov. 2014. +[25] +R. Blumenhagen and E. Plauschinn. Introduction to Conformal Field Theory: With +Applications to String Theory. Lecture Notes in Physics. Springer-Verlag, 2009. url: https: +//www.springer.com/de/book/9783642004490. +[26] +M. Bochicchio. ‘String Field Theory in the Siegel Gauge’. Physics Letters B 188.3 +(Apr. 1987), pp. 330–334. doi: 10.1016/0370-2693(87)91391-8. +[27] +J. B. Bost and T. Jolicoeur. ‘A Holomorphy Property and the Critical Dimension in +String Theory from an Index Theorem’. Physics Letters B 174.3 (July 1986), pp. 273– +276. doi: 10.1016/0370-2693(86)91097-X. +[28] +P. Bouwknegt, J. McCarthy and K. Pilch. ‘BRST Analysis of Physical States for 2D +(Super) Gravity Coupled to (Super) Conformal Matter’. In: New Symmetry Principles +in Quantum Field Theory. Springer, 1992, pp. 413–422. doi: 10.1007/978-1-4615- +3472-3_17. arXiv: hep-th/9110031. +[29] +P. Bouwknegt, J. McCarthy and K. Pilch. ‘BRST analysis of physical states for 2D +gravity coupled to c ≤ 1 matter’. Communications in Mathematical Physics 145.3 +(Apr. 1992), pp. 541–560. doi: 10.1007/BF02099397. +[30] +M. Campbell, P. Nelson and E. Wong. ‘Stress Tensor Perturbations in Conformal +Field Theory’. International Journal of Modern Physics A 06.27 (Nov. 1991), pp. 4909– +4924. doi: 10.1142/S0217751X9100232X. +296 + +[31] +T. Can, M. Laskin and P. Wiegmann. ‘Geometry of Quantum Hall States: Gravita- +tional Anomaly and Kinetic Coefficients’. Annals Phys. 362 (Feb. 2015), pp. 752–794. +doi: 10.1016/j.aop.2015.02.013. arXiv: 1411.3105. +[32] +J. Cardy. ‘Conformal Field Theory and Statistical Mechanics’ (July 2008). arXiv: 0807. +3472. +[33] +R. Catenacci, M. Cornalba, M. Martellini and C. Reina. ‘Algebraic Geometry and +Path Integrals for Closed Strings’. Physics Letters B 172.3 (May 1986), pp. 328–332. +doi: 10.1016/0370-2693(86)90262-5. +[34] +S. Chakrabarti, S. P. Kashyap, B. Sahoo, A. Sen and M. Verma. ‘Subleading Soft +Theorem for Multiple Soft Gravitons’. Journal of High Energy Physics 2017.12 (Dec. +2017). doi: 10.1007/JHEP12(2017)150. arXiv: 1707.06803. +[35] +M. Cho, S. Collier and X. Yin. ‘Strings in Ramond-Ramond Backgrounds from the +Neveu-Schwarz-Ramond Formalism’ (Oct. 2020). arXiv: 1811.00032. +[36] +T. Y. Chow. ‘You Could Have Invented Spectral Sequences’. Notices of the AMS 53.1 +(2006), p. 5. url: http://www.ams.org/notices/200601/fea-chow.pdf. +[37] +J. Collins. ‘A New Approach to the LSZ Reduction Formula’ (Apr. 2019). arXiv: 1904. +10923. +[38] +K. Costello and B. Zwiebach. ‘Hyperbolic String Vertices’ (Aug. 2019). arXiv: 1909. +00033. +[39] +B. Craps and K. Skenderis. ‘Comments on BRST Quantization of Strings’. Journal +of High Energy Physics 2005.05 (May 2005), pp. 001–001. doi: 10.1088/1126- +6708/2005/05/001. arXiv: hep-th/0503038. +[40] +F. David. ‘Conformal Field Theories Coupled to 2-D Gravity in the Conformal +Gauge’. Modern Physics Letters A 03.17 (Dec. 1988), pp. 1651–1656. doi: 10.1142/ +S0217732388001975. +[41] +J. A. de Azcárraga, J. M. Izquierdo and P. K. Townsend. ‘Classical Anomalies of +Supersymmetric Extended Objects’. Physics Letters B 267.3 (Sept. 1991), pp. 366– +373. doi: 10.1016/0370-2693(91)90947-O. +[42] +C. de Lacroix, H. Erbin, S. P. Kashyap, A. Sen and M. Verma. ‘Closed Super- +string Field Theory and Its Applications’. International Journal of Modern Physics A +32.28n29 (Oct. 2017), p. 1730021. doi: 10.1142/S0217751X17300216. arXiv: 1703. +06410. +[43] +C. de Lacroix, H. Erbin and A. Sen. ‘Analyticity and Crossing Symmetry of Super- +string Loop Amplitudes’. Journal of High Energy Physics 2019.5 (May 2019), p. 139. +doi: 10.1007/JHEP05(2019)139. arXiv: 1810.07197. +[44] +B. de Wit, J. Hoppe and H. Nicolai. ‘On the Quantum Mechanics of Supermem- +branes’. Nuclear Physics B 305.4 (Dec. 1988), pp. 545–581. doi: 10.1016/0550- +3213(88)90116-2. +[45] +B. de Wit, M. Lüscher and H. Nicolai. ‘The Supermembrane Is Unstable’. Nuclear +Physics B 320.1 (June 1989), pp. 135–159. doi: 10.1016/0550-3213(89)90214-9. +[46] +G. Delfino and J. Viti. ‘On Three-Point Connectivity in Two-Dimensional Percola- +tion’. Journal of Physics A: Mathematical and Theoretical 44.3 (Jan. 2011), p. 032001. +doi: 10.1088/1751-8113/44/3/032001. arXiv: 1009.1314. +[47] +P. Deligne, P. Etingof, D. S. Freed, L. C. Jeffrey, D. Kazhdan, J. W. Morgan, D. R. +Morrison and E. Witten, eds. Quantum Fields and Strings: A Course for Mathem- +aticians. Volume 1. American Mathematical Society, Oct. 1999. +297 + +[48] +P. Deligne, P. Etingof, D. S. Freed, L. C. Jeffrey, D. Kazhdan, J. W. Morgan, D. R. +Morrison and E. Witten, eds. Quantum Fields and Strings: A Course for Mathem- +aticians. Volume 2. American Mathematical Society, Oct. 1999. +[49] +B. DeWitt. The Global Approach to Quantum Field Theory. 1st edition. Oxford Uni- +versity Press, July 2014. +[50] +E. D’Hoker. ‘Equivalence of Liouville Theory and 2-D Quantum Gravity’. Modern +Physics Letters A 06.09 (Mar. 1991), pp. 745–767. doi: 10.1142/S0217732391000774. +[51] +E. D’Hoker and P. Kurzepa. ‘2D Quantum Gravity and Liouville Theory’. Modern +Physics Letters A 05.18 (July 1990), pp. 1411–1421. doi: 10.1142/S0217732390001608. +[52] +E. D’Hoker and D. H. Phong. ‘Multiloop Amplitudes for the Bosonic Polyakov String’. +Nuclear Physics B 269.1 (May 1986), pp. 205–234. doi: 10.1016/0550-3213(86) +90372-X. +[53] +E. D’Hoker and D. H. Phong. ‘The Geometry of String Perturbation Theory’. Reviews +of Modern Physics 60.4 (Oct. 1988), pp. 917–1065. doi: 10.1103/RevModPhys.60. +917. +[54] +P. Di Francesco, P. Mathieu and D. Senechal. Conformal Field Theory. 2nd edition. +Springer, Jan. 1999. +[55] +J. Distler and H. Kawai. ‘Conformal Field Theory and 2D Quantum Gravity’. Nuclear +Physics B 321.2 (July 1989), pp. 509–527. doi: 10.1016/0550-3213(89)90354-4. +[56] +J. Distler and P. Nelson. ‘Topological Couplings and Contact Terms in 2d Field +Theory’. Communications in Mathematical Physics 138.2 (May 1991), pp. 273–290. +doi: 10.1007/BF02099493. +[57] +J. Distler and P. Nelson. ‘New Discrete States of Strings near a Black Hole’. Nuclear +Physics B 374.1 (Apr. 1992), pp. 123–155. doi: 10.1016/0550-3213(92)90479-U. +[58] +R. Donagi and E. Witten. ‘Supermoduli Space Is Not Projected’ (Apr. 2013). arXiv: 1304. +7798. +[59] +S. Donaldson. Riemann Surfaces. 1st edition. Oxford University Press, May 2011. +[60] +H. Dorn and H.-J. Otto. ‘Two and Three-Point Functions in Liouville Theory’. Nuc- +lear Physics B 429.2 (Oct. 1994), pp. 375–388. doi: 10.1016/0550-3213(94)00352- +1. arXiv: hep-th/9403141. +[61] +H. Dorn and H.-J. Otto. ‘Some Conclusions for Noncritical String Theory Drawn from +Two- and Three-Point Functions in the Liouville Sector’ (Jan. 1995). arXiv: hep- +th/9501019. +[62] +M. J. Duff, T. Inami, C. N. Pope, E. Sezgin and K. S. Stelle. ‘Semiclassical Quant- +ization of the Supermembrane’. Nuclear Physics B 297.3 (Feb. 1988), pp. 515–538. +doi: 10.1016/0550-3213(88)90316-1. +[63] +A. Duncan. The Conceptual Framework of Quantum Field Theory. 1st edition. Oxford +University Press, Oct. 2012. +[64] +H. Erbin. String Field Theory: A Modern Introduction. Lecture Notes in Physics. +Springer, Mar. 2021. doi: 10.1007/978-3-030-65321-7. +[65] +H. Erbin, C. Maccaferri and J. Vošmera. ‘Localization of Effective Actions in Heterotic +String Field Theory’. Journal of High Energy Physics 2020.2 (Feb. 2020), p. 59. +doi: 10.1007/JHEP02(2020)059. arXiv: 1912.05463. +[66] +H. Erbin, J. Maldacena and D. Skliros. ‘Two-Point String Amplitudes’. Journal of +High Energy Physics 2019.7 (June 2019), p. 139. doi: 10.1007/JHEP07(2019)139. +arXiv: 1906.06051. +298 + +[67] +T. Erler. ‘Relating Berkovits and A∞ superstring field theories; large Hilbert space +perspective’. Journal of High Energy Physics 1602 (Oct. 2015), p. 121. doi: 10.1007/ +JHEP02(2016)121. arXiv: 1510.00364. +[68] +T. Erler. ‘Relating Berkovits and A∞ superstring field theories; small Hilbert space +perspective’. Journal of High Energy Physics 1510 (May 2015), p. 157. doi: 10.1007/ +JHEP10(2015)157. arXiv: 1505.02069. +[69] +T. Erler. ‘Superstring Field Theory and the Wess-Zumino-Witten Action’. Journal +of High Energy Physics 2017.10 (Oct. 2017), p. 57. doi: 10.1007/JHEP10(2017)057. +arXiv: 1706.02629. +[70] +T. Erler. ‘Supersymmetry in Open Superstring Field Theory’. Journal of High Energy +Physics 2017.5 (May 2017), p. 113. doi: 10.1007/JHEP05(2017)113. arXiv: 1610. +03251. +[71] +T. Erler. ‘Four Lectures on Closed String Field Theory’. Physics Reports (Jan. 2020), +S0370157320300132. doi: 10.1016/j.physrep.2020.01.003. arXiv: 1905.06785. +[72] +T. Erler. Four Lectures on Analytic Solutions in Open String Field Theory. July 2022. +doi: 10.48550/arXiv.1912.00521. arXiv: 1912.00521. +[73] +T. Erler and S. Konopka. ‘Vertical Integration from the Large Hilbert Space’. Journal +of High Energy Physics 2017.12 (Dec. 2017). doi: 10 . 1007 / JHEP12(2017 ) 112. +arXiv: 1710.07232. +[74] +T. Erler, S. Konopka and I. Sachs. ‘NS-NS Sector of Closed Superstring Field Theory’. +Journal of High Energy Physics 2014.8 (Aug. 2014). doi: 10.1007/JHEP08(2014)158. +arXiv: 1403.0940. +[75] +T. Erler, S. Konopka and I. Sachs. ‘Resolving Witten’s Superstring Field Theory’. +Journal of High Energy Physics 2014.4 (Apr. 2014). doi: 10.1007/JHEP04(2014)150. +arXiv: 1312.2948. +[76] +T. Erler, S. Konopka and I. Sachs. ‘Ramond Equations of Motion in Superstring Field +Theory’. Journal of High Energy Physics 2015.11 (Nov. 2015), p. 199. doi: 10.1007/ +JHEP11(2015)199. arXiv: 1506.05774. +[77] +T. Erler and C. Maccaferri. ‘String Field Theory Solution for Any Open String +Background’. Journal of High Energy Physics 2014.10 (Oct. 2014). doi: 10.1007/ +JHEP10(2014)029. arXiv: 1406.3021. +[78] +T. Erler and C. Maccaferri. ‘String Field Theory Solution for Any Open String Back- +ground. II’. Journal of High Energy Physics 2020.1 (Jan. 2020), p. 21. doi: 10.1007/ +JHEP01(2020)021. arXiv: 1909.11675. +[79] +T. Erler, Y. Okawa and T. Takezaki. ‘A∞ structure from the Berkovits formulation +of open superstring field theory’ (May 2015). arXiv: 1505.01659. +[80] +T. Erler, Y. Okawa and T. Takezaki. ‘Complete Action for Open Superstring Field +Theory with Cyclic A∞ Structure’. Journal of High Energy Physics 2016.8 (Aug. +2016), p. 12. doi: 10.1007/JHEP08(2016)012. arXiv: 1602.02582. +[81] +L. D. Faddeev. ‘Introduction to Functional Methods’. Conf.Proc. C7507281 (1975), +pp. 1–40. doi: 10.1142/9789814412674_0001. +[82] +F. Ferrari and S. Klevtsov. ‘FQHE on Curved Backgrounds, Free Fields and Large +N’. Journal of High Energy Physics 2014.12 (Dec. 2014), p. 086. doi: 10.1007/ +JHEP12(2014)086. arXiv: 1410.6802. +[83] +F. Ferrari, S. Klevtsov and S. Zelditch. ‘Gravitational Actions in Two Dimensions +and the Mabuchi Functional’. Nuclear Physics B 859.3 (June 2012), pp. 341–369. +doi: 10.1016/j.nuclphysb.2012.02.003. arXiv: 1112.1352. +299 + +[84] +M. Flohr. ‘On Modular Invariant Partition Functions of Conformal Field Theories +with Logarithmic Operators’. International Journal of Modern Physics A 11.22 (Sept. +1996), pp. 4147–4172. doi: 10.1142/S0217751X96001954. arXiv: hep-th/9509166. +[85] +M. Flohr. ‘Bits and Pieces in Logarithmic Conformal Field Theory’. International +Journal of Modern Physics A 18.25 (Oct. 2003), pp. 4497–4591. doi: 10 . 1142 / +S0217751X03016859. arXiv: hep-th/0111228. +[86] +D. Z. Freedman and A. Van Proeyen. Supergravity. Cambridge University Press, May +2012. +[87] +K. Fujikawa, U. Lindström, N. K. Nielsen, M. Rocek and P. van Nieuwenhuizen. ‘Reg- +ularized BRST-coordinate-invariant Measure’. Physical Review D 37.2 (Jan. 1988), +pp. 391–405. doi: 10.1103/PhysRevD.37.391. +[88] +A. Fuster, M. Henneaux and A. Maas. ‘BRST-antifield Quantization: A Short Re- +view’. International Journal of Geometric Methods in Modern Physics 2 (June 2005), +pp. 939–964. doi: 10.1142/S0219887805000892. arXiv: hep-th/0506098. +[89] +M. R. Gaberdiel. ‘An Introduction to Conformal Field Theory’ (Oct. 1999). doi: 10. +1088/0034-4885/63/4/203. arXiv: hep-th/9910156. +[90] +M. R. Gaberdiel. ‘An Algebraic Approach to Logarithmic Conformal Field The- +ory’. International Journal of Modern Physics A 18.25 (Oct. 2003), pp. 4593–4638. +doi: 10.1142/S0217751X03016860. arXiv: hep-th/0111260. +[91] +P. Ginsparg. ‘Applied Conformal Field Theory’ (Nov. 1988). arXiv: hep-th/9108028. +[92] +P. Goddard, J. Goldstone, C. Rebbi and C. Thorn. ‘Quantum Dynamics of a Massless +Relativistic String’. Nuclear Physics B 56.1 (May 1973), pp. 109–135. doi: 10.1016/ +0550-3213(73)90223-X. +[93] +J. Gomis, J. Paris and S. Samuel. ‘Antibracket, Antifields and Gauge-Theory Quant- +ization’. Physics Reports 259.1-2 (Aug. 1995), pp. 1–145. doi: 10 . 1016 / 0370 - +1573(94)00112-G. arXiv: hep-th/9412228. +[94] +K. Goto and H. Kunitomo. ‘Construction of Action for Heterotic String Field Theory +Including the Ramond Sector’ (June 2016). arXiv: 1606.07194. +[95] +M. B. Green, J. H. Schwarz and E. Witten. Superstring Theory: Introduction. Vol. 1. +Cambridge University Press, July 1988. +[96] +E. Guadagnini. ‘Central Charge, Trace and Gravitational Anomalies in Two Dimen- +sions’. Physical Review D 38.8 (Oct. 1988), pp. 2482–2489. doi: 10.1103/PhysRevD. +38.2482. +[97] +V. Gurarie. ‘Logarithmic Operators in Conformal Field Theory’. Nuclear Physics B +410.3 (Dec. 1993), pp. 535–549. doi: 10.1016/0550-3213(93)90528-W. arXiv: hep- +th/9303160. +[98] +H. W. Hamber. Quantum Gravitation: The Feynman Path Integral Approach. Springer- +Verlag, 2009. +[99] +D. Harlow, J. Maltz and E. Witten. ‘Analytic Continuation of Liouville Theory’. +Journal of High Energy Physics 12 (2011), p. 071. doi: 10.1007/JHEP12(2011)071. +arXiv: 1108.4417. +[100] +B. Hatfield. Quantum Field Theory of Point Particles and Strings. Addison Wesley, +Apr. 1998. +[101] +M. Hatsuda, P. van Nieuwenhuizen, W. Troost and A. Van Proeyen. ‘The Regularized +Phase Space Path Integral Measure for a Scalar Field Coupled to Gravity’. Nuclear +Physics B 335.1 (Apr. 1990), pp. 166–196. doi: 10.1016/0550-3213(90)90176-E. +300 + +[102] +M. Headrick and B. Zwiebach. ‘Convex Programs for Minimal-Area Problems’. Com- +munications in Mathematical Physics 377.3 (2020), pp. 2217–2285. doi: 10.1007/ +s00220-020-03732-1. arXiv: 1806.00449. +[103] +M. Headrick and B. Zwiebach. ‘Minimal-Area Metrics on the Swiss Cross and Punc- +tured Torus’. Communications in Mathematical Physics (Mar. 2020). doi: 10.1007/ +s00220-020-03734-z. arXiv: 1806.00450. +[104] +M. Henkel. Conformal Invariance and Critical Phenomena. 1st edition. Springer, Dec. +2010. +[105] +M. Henneaux and C. Teitelboim. Quantization of Gauge Systems. Princeton Univer- +sity Press, Aug. 1994. +[106] +O. Hohm, V. Kupriyanov, D. Lüst and M. Traube. ‘General constructions of L∞ +algebras’ (Sept. 2017). arXiv: 1709.10004. +[107] +O. Hohm, D. Lüst and B. Zwiebach. ‘The Spacetime of Double Field Theory: Re- +view, Remarks, and Outlook’. Fortschritte der Physik 61.10 (Oct. 2013), pp. 926–966. +doi: 10.1002/prop.201300024. arXiv: 1309.2977. +[108] +O. Hohm and B. Zwiebach. ‘L∞ Algebras and Field Theory’. Fortschritte der Physik +65.3-4 (2017), p. 1700014. doi: 10.1002/prop.201700014. arXiv: 1701.08824. +[109] +C. M. Hull. ‘Doubled Geometry and T-Folds’. Journal of High Energy Physics 2007.07 +(July 2007), pp. 080–080. doi: 10.1088/1126- 6708/2007/07/080. arXiv: hep- +th/0605149. +[110] +C. Hull and B. Zwiebach. ‘Double Field Theory’. Journal of High Energy Phys- +ics 2009.09 (Sept. 2009), pp. 099–099. doi: 10.1088/1126- 6708/2009/09/099. +arXiv: 0904.4664. +[111] +Y. Iimori, T. Noumi, Y. Okawa and S. Torii. ‘From the Berkovits Formulation to +the Witten Formulation in Open Superstring Field Theory’. Journal of High Energy +Physics 2014.3 (Mar. 2014), p. 44. doi: 10.1007/JHEP03(2014)044. arXiv: 1312. +1677. +[112] +Y. Ikhlef, J. L. Jacobsen and H. Saleur. ‘Three-point functions in c ≤ 1 Liouville +theory and conformal loop ensembles (v1)’. Physical Review Letters 116.13 (Mar. +2016). doi: 10.1103/PhysRevLett.116.130601. arXiv: 1509.03538. +[113] +A. Iorio, L. O’Raifeartaigh, I. Sachs and C. Wiesendanger. ‘Weyl-Gauging and Con- +formal Invariance’. Nuclear Physics B 495.1-2 (June 1997), pp. 433–450. doi: 10. +1016/S0550-3213(97)00190-9. arXiv: hep-th/9607110. +[114] +N. Ishibashi. ‘Multiloop Amplitudes of Light-cone Gauge String Field Theory for +Type II Superstrings’ (Oct. 2018). +[115] +N. Ishibashi and K. Murakami. ‘Multiloop Amplitudes of Light-cone Gauge Bosonic +String Field Theory in Noncritical Dimensions’. Journal of High Energy Physics +2013.9 (Sept. 2013). doi: 10.1007/JHEP09(2013)053. arXiv: 1307.6001. +[116] +N. Ishibashi and K. Murakami. ‘Worldsheet Theory of Light-Cone Gauge Noncritical +Strings on Higher Genus Riemann Surfaces’. Journal of High Energy Physics 2016.6 +(June 2016). doi: 10.1007/JHEP06(2016)087. arXiv: 1603.08337. +[117] +N. Ishibashi and K. Murakami. ‘Multiloop Amplitudes of Light-cone Gauge Super- +string Field Theory: Odd Spin Structure Contributions’. Journal of High Energy +Physics 2018.3 (Mar. 2018), p. 63. doi: 10.1007/JHEP03(2018)063. arXiv: 1712. +09049. +[118] +K. Itoh. ‘BRST Quantization of Polyakov’s Two-Dimensional Gravity’. Nuclear Phys- +ics B 342.2 (Oct. 1990), pp. 449–470. doi: 10.1016/0550-3213(90)90198-M. +301 + +[119] +K. Itoh and N. Ohta. ‘Spectrum of Two-Dimensional (Super)Gravity’. Progress of +Theoretical Physics Supplement 110 (1992), pp. 97–116. doi: 10.1143/PTPS.110.97. +arXiv: hep-th/9201034. +[120] +C. Itzykson and J.-M. Drouffe. Théorie statistique des champs. Volume 2. EDP Sci- +ences, 2000. +[121] +C. Itzykson and J.-B. Zuber. Quantum Field Theory. Dover Publications, Feb. 2006. +[122] +R. Jackiw. ‘Another View on Massless Matter-Gravity Fields in Two Dimensions’ +(Jan. 1995). arXiv: hep-th/9501016. +[123] +C. V. Johnson. D-Branes. Cambridge University Press, Nov. 2006. +[124] +M. Kaku. Introduction to Superstrings and M-Theory. Springer, 1999. +[125] +M. Kaku. Strings, Conformal Fields, and M-Theory. Springer, 1999. +[126] +S. V. Ketov. Conformal Field Theory. World Scientific, Feb. 1995. +[127] +K. Kikkawa and M. Yamasaki. ‘Can the Membrane Be a Unification Model?’ Progress +of Theoretical Physics 76.6 (Dec. 1986), pp. 1379–1389. doi: 10.1143/PTP.76.1379. +[128] +E. Kiritsis. String Theory in a Nutshell. Princeton University Press, 2007. +[129] +M. Knecht, S. Lazzarini and F. Thuillier. ‘Shifting the Weyl Anomaly to the Chirally +Split Diffeomorphism Anomaly in Two Dimensions’. Physics Letters B 251.2 (Nov. +1990), pp. 279–283. doi: 10.1016/0370-2693(90)90936-Z. +[130] +I. I. Kogan and N. E. Mavromatos. ‘World-Sheet Logarithmic Operators and Target +Space Symmetries in String Theory’. Physics Letters B 375.1-4 (May 1996), pp. 111– +120. doi: 10.1016/0370-2693(96)00195-5. arXiv: hep-th/9512210. +[131] +S. Konopka and I. Sachs. ‘Open Superstring Field Theory on the Restricted Hilbert +Space’. Journal of High Energy Physics 2016.4 (Apr. 2016), pp. 1–12. doi: 10.1007/ +JHEP04(2016)164. arXiv: 1602.02583. +[132] +M. Kroyter. ‘Superstring Field Theory in the Democratic Picture’. Adv.Theor.Math.Phys. +15.MIT-CTP-4037 (2009), pp. 741–781. doi: 10.4310/ATMP.2011.v15.n3.a3. +arXiv: 0911.2962. +[133] +M. Kroyter. ‘Democratic Superstring Field Theory: Gauge Fixing’. Journal of High +Energy Physics 2011.3 (Mar. 2011). doi: 10.1007/JHEP03(2011)081. arXiv: 1010. +1662. +[134] +M. Kroyter, Y. Okawa, M. Schnabl, S. Torii and B. Zwiebach. ‘Open Superstring Field +Theory I: Gauge Fixing, Ghost Structure, and Propagator’. Journal of High Energy +Physics 2012.3 (Mar. 2012). doi: 10.1007/JHEP03(2012)030. arXiv: 1201.1761. +[135] +M. Kudrna and C. Maccaferri. ‘BCFT Moduli Space in Level Truncation’. Journal +of High Energy Physics 2016.4 (Apr. 2016). doi: 10 . 1007 / JHEP04(2016 ) 057. +arXiv: 1601.04046. +[136] +M. Kudrna, T. Masuda, Y. Okawa, M. Schnabl and K. Yoshida. ‘Gauge-Invariant Ob- +servables and Marginal Deformations in Open String Field Theory’. Journal of High +Energy Physics 2013.1 (Jan. 2013). doi: 10.1007/JHEP01(2013)103. arXiv: 1207. +3335. +[137] +M. Kudrna and M. Schnabl. ‘Universal Solutions in Open String Field Theory’. +arxiv:1812.03221 (Dec. 2018). arXiv: 1812.03221. +[138] +T. Kugo, H. Kunitomo and K. Suehiro. ‘Non-Polynomial Closed String Field Theory’. +Physics Letters B 226.1 (Aug. 1989), pp. 48–54. doi: 10.1016/0370-2693(89)90287- +6. +302 + +[139] +T. Kugo and K. Suehiro. ‘Nonpolynomial Closed String Field Theory: Action and +Its Gauge Invariance’. Nuclear Physics B 337.2 (June 1990), pp. 434–466. doi: 10. +1016/0550-3213(90)90277-K. +[140] +H. Kunitomo. ‘First-Order Equations of Motion for Heterotic String Field The- +ory’. Progress of Theoretical and Experimental Physics 2014.9 (Sept. 2014), 93B07–0. +doi: 10.1093/ptep/ptu125. arXiv: 1407.0801. +[141] +H. Kunitomo. ‘The Ramond Sector of Heterotic String Field Theory’. Progress of +Theoretical and Experimental Physics 2014.4 (Apr. 2014), 43B01–0. doi: 10.1093/ +ptep/ptu032. arXiv: 1312.7197. +[142] +H. Kunitomo. ‘Symmetries and Feynman Rules for Ramond Sector in Heterotic String +Field Theory’. Progress of Theoretical and Experimental Physics 2015.9 (Sept. 2015), +093B02. doi: 10.1093/ptep/ptv117. arXiv: 1506.08926. +[143] +H. Kunitomo. ‘Space-Time Supersymmetry in WZW-like Open Superstring Field +Theory’. Progress of Theoretical and Experimental Physics 2017.4 (Apr. 2017). doi: 10. +1093/ptep/ptx028. arXiv: 1612.08508. +[144] +H. Kunitomo and Y. Okawa. ‘Complete Action for Open Superstring Field The- +ory’. Progress of Theoretical and Experimental Physics 2016.2 (Feb. 2016), 023B01. +doi: 10.1093/ptep/ptv189. arXiv: 1508.00366. +[145] +H. Kunitomo and T. Sugimoto. ‘Heterotic String Field Theory with Cyclic L-infinity +Structure’. Progress of Theoretical and Experimental Physics 2019.6 (June 2019), +063B02. doi: 10.1093/ptep/ptz051. arXiv: 1902.02991. +[146] +H. Kunitomo and T. Sugimoto. ‘Type II Superstring Field Theory with Cyclic L- +infinity Structure’ (Mar. 2020). arXiv: 1911.04103. +[147] +J. M. F. Labastida and M. Pernici. ‘BRST Quantization in the Siegel Gauge’. Physics +Letters B 194.4 (Aug. 1987), pp. 511–517. doi: 10.1016/0370-2693(87)90226-7. +[148] +T. Lada and M. Markl. ‘Strongly Homotopy Lie Algebras’ (June 1994). arXiv: hep- +th/9406095. +[149] +T. Lada and J. Stasheff. ‘Introduction to Sh Lie Algebras for Physicists’. Interna- +tional Journal of Theoretical Physics 32.7 (July 1993), pp. 1087–1103. doi: 10.1007/ +BF00671791. arXiv: hep-th/9209099. +[150] +A. Laddha and A. Sen. ‘Sub-Subleading Soft Graviton Theorem in Generic Theories +of Quantum Gravity’. Journal of High Energy Physics 2017.10 (Oct. 2017), p. 065. +doi: 10.1007/JHEP10(2017)065. arXiv: 1706.00759. +[151] +I. D. Lawrie. A Unified Grand Tour of Theoretical Physics. 3rd edition. CRC Press, +Nov. 2012. +[152] +H. Luckock and I. Moss. ‘The Quantum Geometry of Random Surfaces and Spin- +ning Membranes’. Classical and Quantum Gravity 6.12 (Dec. 1989), pp. 1993–2027. +doi: 10.1088/0264-9381/6/12/025. +[153] +C. Maccaferri and A. Merlano. ‘Localization of Effective Actions in Open Superstring +Field Theory’. Journal of High Energy Physics 2018.3 (Mar. 2018), p. 112. doi: 10. +1007/JHEP03(2018)112. arXiv: 1801.07607. +[154] +C. Maccaferri and A. Merlano. ‘Localization of Effective Actions in Open Superstring +Field Theory: Small Hilbert Space’. Journal of High Energy Physics 1906.arXiv:1905.04958 +(May 2019), p. 101. doi: 10.1007/JHEP06(2019)101. arXiv: 1905.04958. +[155] +P. Mansfield. ‘Nilpotent BRST Invariance of the Interacting Polyakov String’. Nuclear +Physics B 283.Supplement C (Jan. 1987), pp. 551–576. doi: 10.1016/0550-3213(87) +90286-0. +303 + +[156] +U. Marquard, R. Kaiser and M. Scholl. ‘Lorentz Algebra and Critical Dimension for +the Supermembrane’. Physics Letters B 227.2 (Aug. 1989), pp. 234–238. doi: 10. +1016/S0370-2693(89)80028-0. +[157] +U. Marquard and M. Scholl. ‘Conditions of the Embedding Space of P-Branes from +Their Constraint Algebras’. Physics Letters B 209.4 (Aug. 1988), pp. 434–440. doi: 10. +1016/0370-2693(88)91169-0. +[158] +U. Marquard and M. Scholl. ‘Lorentz Algebra and Critical Dimension for the Bosonic +Membrane’. Physics Letters B 227.2 (Aug. 1989), pp. 227–233. doi: 10.1016/S0370- +2693(89)80027-9. +[159] +H. Matsunaga. ‘Comments on Complete Actions for Open Superstring Field Theory’. +Journal of High Energy Physics 2016.11 (Nov. 2016). doi: 10.1007/JHEP11(2016) +115. arXiv: 1510.06023. +[160] +N. Mavromatos and J. Miramontes. ‘Regularizing the Functional Integral in 2D- +Quantum Gravity’. Modern Physics Letters A 04.19 (Sept. 1989), pp. 1847–1853. +doi: 10.1142/S0217732389002082. +[161] +G. Moore and P. Nelson. ‘Measure for Moduli The Polyakov String Has No Nonlocal +Anomalies’. Nuclear Physics B 266.1 (Mar. 1986), pp. 58–74. doi: 10.1016/0550- +3213(86)90177-X. +[162] +S. F. Moosavian and R. Pius. ‘Hyperbolic Geometry of Superstring Perturbation +Theory (V2)’ (Mar. 2017). arXiv: 1703.10563. +[163] +S. F. Moosavian and R. Pius. ‘Hyperbolic Geometry and Closed Bosonic String Field +Theory I: The String Vertices Via Hyperbolic Riemann Surfaces’. Journal of High En- +ergy Physics 1908.arXiv:1706.07366 (Aug. 2019), p. 157. doi: 10.1007/JHEP08(2019) +157. arXiv: 1706.07366. +[164] +S. F. Moosavian and R. Pius. ‘Hyperbolic Geometry and Closed Bosonic String Field +Theory II: The Rules for Evaluating the Quantum BV Master Action’. Journal of +High Energy Physics 1908.arXiv:1708.04977 (Aug. 2019), p. 177. doi: 10 . 1007 / +JHEP08(2019)177. arXiv: 1708.04977. +[165] +S. F. Moosavian, A. Sen and M. Verma. ‘Superstring Field Theory with Open and +Closed Strings’. Journal of High Energy Physics 2001 (Jan. 2020), p. 183. doi: 10. +1007/JHEP01(2020)183. arXiv: 1907.10632. +[166] +S. F. Moosavian and Y. Zhou. ‘On the Existence of Heterotic-String and Type-II- +Superstring Field Theory Vertices’ (Nov. 2019). arXiv: 1911.04343. +[167] +K. Muenster and I. Sachs. ‘On Homotopy Algebras and Quantum String Field Theory’ +(Mar. 2013). arXiv: 1303.3444. +[168] +K. Muenster and I. Sachs. ‘Quantum Open-Closed Homotopy Algebra and String +Field Theory’. Communications in Mathematical Physics 321.3 (Aug. 2013), pp. 769– +801. doi: 10.1007/s00220-012-1654-1. arXiv: 1109.4101. +[169] +K. Muenster and I. Sachs. ‘Homotopy Classification of Bosonic String Field Theory’. +Communications in Mathematical Physics 330 (2014), pp. 1227–1262. doi: 10.1007/ +s00220-014-2027-8. arXiv: 1208.5626. +[170] +S. Mukhi. ‘Extra States in C<1 String Theory’ (Nov. 1991). arXiv: hep-th/9111013. +[171] +G. Mussardo. Statistical Field Theory: An Introduction to Exactly Solved Models in +Statistical Physics. Oxford University Press, 2009. +[172] +M. Nakahara. Geometry, Topology and Physics. 2nd edition. Institute of Physics +Publishing, June 2003. +304 + +[173] +P. Nelson. ‘Lectures on Strings and Moduli Space’. Physics Reports 149.6 (May 1987), +pp. 337–375. doi: 10.1016/0370-1573(87)90082-2. +[174] +P. Nelson. ‘Covariant Insertion of General Vertex Operators’. Physical Review Letters +62.9 (Feb. 1989), pp. 993–996. doi: 10.1103/PhysRevLett.62.993. +[175] +K. Ohmori and Y. Okawa. ‘Open Superstring Field Theory Based on the Supermoduli +Space’. Journal of High Energy Physics 2018.4 (Apr. 2018), p. 35. doi: 10.1007/ +JHEP04(2018)035. arXiv: 1703.08214. +[176] +K. Ohmori. ‘A Review on Tachyon Condensation in Open String Field Theories’ (Feb. +2001). arXiv: hep-th/0102085. +[177] +N. Ohta. ‘Discrete States in Two-Dimensional Quantum Gravity’ (June 1992). arXiv: hep- +th/9206012. +[178] +Y. Okawa. Construction of Superstring Field Theories. Feb. 2018. url: http://www. +hri.res.in/~strings/okawa_school.pdf. +[179] +Y. Okawa and R. Sakaguchi. ‘Closed String Field Theory without the Level-Matching +Condition (Draft)’ (Jan. 2018). +[180] +Y. Okawa and B. Zwiebach. ‘Heterotic String Field Theory’. Journal of High Energy +Physics 2004.07 (July 2004), pp. 042–042. doi: 10.1088/1126-6708/2004/07/042. +arXiv: hep-th/0406212. +[181] +F. Paccanoni, P. Pasti and M. Tonin. ‘Some Remarks on the Consistency of Quantum +Supermembranes’. Modern Physics Letters A (1989). doi: 10.1142/S0217732389000940. +[182] +M. Picco, R. Santachiara, J. Viti and G. Delfino. ‘Connectivities of Potts Fortuin- +Kasteleyn Clusters and Time-like Liouville Correlator’. Nuclear Physics B 875.3 (Oct. +2013), pp. 719–737. doi: 10.1016/j.nuclphysb.2013.07.014. arXiv: 1304.6511. +[183] +R. Pius. ‘Quantum Closed Superstring Field Theory and Hyperbolic Geometry I: +Construction of String Vertices’ (Aug. 2018). arXiv: 1808.09441. +[184] +R. Pius, A. Rudra and A. Sen. ‘Mass Renormalization in String Theory: General +States’. Journal of High Energy Physics 2014.7 (July 2014). doi: 10.1007/JHEP07(2014) +062. arXiv: 1401.7014. +[185] +R. Pius, A. Rudra and A. Sen. ‘Mass Renormalization in String Theory: Special +States’. Journal of High Energy Physics 2014.7 (July 2014). doi: 10.1007/JHEP07(2014) +058. arXiv: 1311.1257. +[186] +R. Pius, A. Rudra and A. Sen. ‘String Perturbation Theory Around Dynamically +Shifted Vacuum’. Journal of High Energy Physics 2014.10 (Oct. 2014). doi: 10. +1007/JHEP10(2014)070. arXiv: 1404.6254. +[187] +R. Pius and A. Sen. ‘Cutkosky Rules for Superstring Field Theory’. Journal of High +Energy Physics 2016.10 (Oct. 2016). doi: 10.1007/JHEP10(2016)024. arXiv: 1604. +01783. +[188] +R. Pius and A. Sen. ‘Unitarity of the Box Diagram’. Journal of High Energy Physics +2018.11 (Nov. 2018), p. 094. doi: 10.1007/JHEP11(2018)094. arXiv: 1805.00984. +[189] +E. Plauschinn. ‘Non-Geometric Backgrounds in String Theory’ (Nov. 2018). +[190] +E. Poisson. A Relativist’s Toolkit: The Mathematics of Black-Hole Mechanics. 1st +edition. Cambridge University Press, Nov. 2007. +[191] +J. Polchinski. ‘Evaluation of the One Loop String Path Integral’. Communications in +Mathematical Physics 104.1 (Mar. 1986), pp. 37–47. doi: 10.1007/BF01210791. +[192] +J. Polchinski. ‘What Is String Theory?’ (Nov. 1994). arXiv: hep-th/9411028. +305 + +[193] +J. Polchinski. String Theory: Volume 1, An Introduction to the Bosonic String. Cam- +bridge University Press, June 2005. +[194] +J. Polchinski. String Theory, Volume 2: Superstring Theory and Beyond. Cambridge +University Press, July 2005. +[195] +J. Preskill. ‘Gauge Anomalies in an Effective Field Theory’. Annals of Physics 210.2 +(Sept. 1991), pp. 323–379. doi: 10.1016/0003-4916(91)90046-B. +[196] +J. D. Qualls. ‘Lectures on Conformal Field Theory’ (Nov. 2015). arXiv: 1511.04074. +[197] +S. Rahman and B. Zwiebach. ‘Vacuum Vertices and the Ghost-Dilaton’. Nuclear +Physics B 471.1-2 (July 1996), pp. 233–245. doi: 10.1016/0550-3213(96)00179-4. +arXiv: hep-th/9507038. +[198] +K. Ranganathan. ‘Nearby CFT’s in the Operator Formalism: The Role of a Con- +nection’. Nuclear Physics B 408.1 (Nov. 1993), pp. 180–206. doi: 10.1016/0550- +3213(93)90136-D. arXiv: hep-th/9210090. +[199] +K. Ranganathan, H. Sonoda and B. Zwiebach. ‘Connections on the State-Space over +Conformal Field Theories’. Nuclear Physics B 414.1-2 (Feb. 1994), pp. 405–460. +doi: 10.1016/0550-3213(94)90436-7. arXiv: hep-th/9304053. +[200] +S. Ribault. ‘Conformal Field Theory on the Plane (V5)’ (June 2014). arXiv: 1406. +4290. +[201] +S. Ribault. ‘Minimal Lectures on Two-Dimensional Conformal Field Theory’. SciPost +Physics Lecture Notes (June 2018). doi: 10 . 21468 / SciPostPhysLectNotes . 1. +arXiv: 1609.09523. +[202] +S. Ribault and R. Santachiara. ‘Liouville Theory with a Central Charge Less than +One’. Journal of High Energy Physics 2015.8 (Aug. 2015), p. 109. doi: 10.1007/ +JHEP08(2015)109. arXiv: 1503.02067. +[203] +S. Rychkov. ‘EPFL Lectures on Conformal Field Theory in D>= 3 Dimensions’ (Jan. +2016). arXiv: 1601.05000. +[204] +A. N. Schellekens. Introduction to Conformal Field Theory. 2016. url: http://www. +nikhef.nl/~t58/CFT.pdf. +[205] +V. Schomerus. A Primer on String Theory. 1st edition. Cambridge University Press, +June 2017. +[206] +M. Schottenloher. A Mathematical Introduction to Conformal Field Theory. 2nd edi- +tion. Springer, Nov. 2008. +[207] +N. Seiberg. ‘Notes on Quantum Liouville Theory and Quantum Gravity’. Progress of +Theoretical Physics Supplement 102 (1990), pp. 319–349. doi: 10.1143/PTPS.102. +319. +[208] +S. Seki and T. Takahashi. ‘Two-Point String Amplitudes Revisited by Operator Form- +alism’. Physics Letters B 800 (Jan. 2020), p. 135078. doi: 10.1016/j.physletb. +2019.135078. arXiv: 1909.03672. +[209] +A. Sen. ‘On the Background Independence of String Field Theory: II. Analysis of on- +Shell S-matrix Elements’. Nuclear Physics B 347.1 (Dec. 1990), pp. 270–318. doi: 10. +1016/0550-3213(90)90560-Z. +[210] +A. Sen. ‘On the Background Independence of String Field Theory’. Nuclear Physics +B 345.2–3 (Dec. 1990), pp. 551–583. doi: 10.1016/0550-3213(90)90400-8. +[211] +A. Sen. ‘On the Background Independence of String Field Theory: III. Explicit Field +Redefinitions’. Nuclear Physics B 391.3 (Feb. 1993), pp. 550–590. doi: 10.1016/ +0550-3213(93)90084-3. arXiv: hep-th/9201041. +306 + +[212] +A. Sen. String Theory 1. 2011. url: http://www.hri.res.in/~sen/combind.pdf. +[213] +A. Sen. ‘Gauge Invariant 1PI Effective Action for Superstring Field Theory’. Journal +of High Energy Physics 1506 (June 2015), p. 022. doi: 10.1007/JHEP06(2015)022. +arXiv: 1411.7478. +[214] +A. Sen. ‘Gauge Invariant 1PI Effective Superstring Field Theory: Inclusion of the +Ramond Sector’. Journal of High Energy Physics 2015.8 (Aug. 2015). doi: 10.1007/ +JHEP08(2015)025. arXiv: 1501.00988. +[215] +A. Sen. ‘Off-Shell Amplitudes in Superstring Theory’. Fortschritte der Physik 63.3-4 +(Apr. 2015), pp. 149–188. doi: 10.1002/prop.201500002. arXiv: 1408.0571. +[216] +A. Sen. ‘Supersymmetry Restoration in Superstring Perturbation Theory’. Journal +of High Energy Physics 2015.12 (Dec. 2015). doi: 10 . 1007 / JHEP12(2015 ) 075. +arXiv: 1508.02481. +[217] +A. Sen. ‘Ultraviolet and Infrared Divergences in Superstring Theory’ (Nov. 2015). +arXiv: 1512.00026. +[218] +A. Sen. ‘BV Master Action for Heterotic and Type II String Field Theories’. Journal +of High Energy Physics 2016.2 (Feb. 2016). doi: 10.1007/JHEP02(2016)087. arXiv: 1508. +05387. +[219] +A. Sen. ‘One Loop Mass Renormalization of Unstable Particles in Superstring The- +ory’. Journal of High Energy Physics 2016.11 (Nov. 2016). doi: 10.1007/JHEP11(2016) +050. arXiv: 1607.06500. +[220] +A. Sen. ‘Reality of Superstring Field Theory Action’. Journal of High Energy Physics +2016.11 (Nov. 2016), p. 014. doi: 10.1007/JHEP11(2016)014. arXiv: 1606.03455. +[221] +A. Sen. ‘Unitarity of Superstring Field Theory’. Journal of High Energy Physics +2016.12 (Dec. 2016). doi: 10.1007/JHEP12(2016)115. arXiv: 1607.08244. +[222] +A. Sen. ‘Equivalence of Two Contour Prescriptions in Superstring Perturbation The- +ory’. Journal of High Energy Physics 2017.04 (Apr. 2017). doi: 10.1007/JHEP04(2017) +025. arXiv: 1610.00443. +[223] +A. Sen. ‘Soft Theorems in Superstring Theory’. Journal of High Energy Physics +2017.06 (June 2017). doi: 10.1007/JHEP06(2017)113. arXiv: 1702.03934. +[224] +A. Sen. ‘Subleading Soft Graviton Theorem for Loop Amplitudes’. Journal of High +Energy Physics 2017.11 (Nov. 2017). doi: 10.1007/JHEP11(2017)123. arXiv: 1703. +00024. +[225] +A. Sen. ‘Wilsonian Effective Action of Superstring Theory’. Journal of High Energy +Physics 2017.1 (Jan. 2017), p. 108. doi: 10.1007/JHEP01(2017)108. arXiv: 1609. +00459. +[226] +A. Sen. ‘Background Independence of Closed Superstring Field Theory’. Journal of +High Energy Physics 2018.2 (Feb. 2018), p. 155. doi: 10.1007/JHEP02(2018)155. +arXiv: 1711.08468. +[227] +A. Sen. ‘String Field Theory as World-sheet UV Regulator’. Journal of High Energy +Physics 2019.10 (Oct. 2019), p. 119. doi: 10.1007/JHEP10(2019)119. arXiv: 1902. +00263. +[228] +A. Sen. ‘D-Instanton Perturbation Theory’ (May 2020). arXiv: 2002.04043. +[229] +A. Sen. ‘Fixing an Ambiguity in Two Dimensional String Theory Using String Field +Theory’. Journal of High Energy Physics 2020.3 (Mar. 2020), p. 5. doi: 10.1007/ +JHEP03(2020)005. arXiv: 1908.02782. +307 + +[230] +A. Sen and E. Witten. ‘Filling the Gaps with PCO’s’. Journal of High Energy Phys- +ics 1509.arXiv:1504.00609 (Apr. 2015), p. 004. doi: 10.1007/JHEP09(2015)004. +arXiv: 1504.00609. +[231] +A. Sen and B. Zwiebach. ‘A Note on Gauge Transformations in Batalin-Vilkovisky +Theory’. Physics Letters B 320.1-2 (Jan. 1994), pp. 29–35. doi: 10.1016/0370- +2693(94)90819-2. arXiv: hep-th/9309027. +[232] +A. Sen and B. Zwiebach. ‘A Proof of Local Background Independence of Classical +Closed String Field Theory’. Nuclear Physics B 414.3 (Feb. 1994), pp. 649–711. +doi: 10.1016/0550-3213(94)90258-5. arXiv: hep-th/9307088. +[233] +A. Sen and B. Zwiebach. ‘Quantum Background Independence of Closed String Field +Theory’. Nuclear Physics B 423.2-3 (July 1994), pp. 580–630. doi: 10.1016/0550- +3213(94)90145-7. arXiv: hep-th/9311009. +[234] +A. Sen and B. Zwiebach. ‘Background Independent Algebraic Structures in Closed +String Field Theory’. Communications in Mathematical Physics 177.2 (Apr. 1996), +pp. 305–326. doi: 10.1007/BF02101895. arXiv: hep-th/9408053. +[235] +W. Siegel. ‘Introduction to String Field Theory’ (July 2001). +[236] +D. Simmons-Duffin. ‘TASI Lectures on the Conformal Bootstrap’ (Feb. 2016). arXiv: 1602. +07982. +[237] +D. P. Skliros, E. J. Copeland and P. M. Saffin. ‘Highly Excited Strings I: Generating +Function’. Nuclear Physics B 916 (2017), pp. 143–207. doi: 10.1016/j.nuclphysb. +2016.12.022. arXiv: 1611.06498. +[238] +H. Sonoda. ‘Connection on the Theory Space’. In: arXiv:Hep-Th/9306119. June 1993. +arXiv: hep-th/9306119. +[239] +M. Srednicki. Quantum Field Theory. 1st edition. Cambridge University Press, Feb. +2007. +[240] +G. ’t Hooft. Introduction to String Theory. May 2004. url: https://www.staff. +science.uu.nl/~hooft101/lectures/stringnotes.pdf. +[241] +T. Takezaki. ‘Open Superstring Field Theory Including the Ramond Sector Based on +the Supermoduli Space’ (Jan. 2019). arXiv: 1901.02176. +[242] +W. Taylor and B. Zwiebach. ‘D-Branes, Tachyons, and String Field Theory’. Strings, +Branes and Extra Dimensions (Mar. 2004), pp. 641–760. doi: 10.1142/9789812702821_ +0012. arXiv: hep-th/0311017. +[243] +J. Teschner. ‘A Guide to Two-Dimensional Conformal Field Theory’. Les Houches +Lect.Notes 106.arXiv:1708.00680 (Aug. 2017). doi: 10.1093/oso/9780198828150. +003.0002. arXiv: 1708.00680. +[244] +C. B. Thorn. ‘Perturbation Theory for Quantized String Fields’. Nuclear Physics B +287.Supplement C (Jan. 1987), pp. 61–92. doi: 10.1016/0550-3213(87)90096-4. +[245] +C. B. Thorn. ‘String Field Theory’. Physics Reports 175.1–2 (Apr. 1989), pp. 1–101. +doi: 10.1016/0370-1573(89)90015-X. +[246] +D. Tong. ‘Lectures on String Theory’ (Aug. 2009). arXiv: 0908.0333. +[247] +J. W. van Holten. ‘Aspects of BRST Quantization’. 659 (Dec. 2004), pp. 99–166. +doi: 10.1007/978-3-540-31532-2_3. arXiv: hep-th/0201124. +[248] +J. Vošmera. ‘Generalized ADHM Equations from Marginal Deformations in Open Su- +perstring Field Theory’. Journal of High Energy Physics 2019.12 (Dec. 2019), p. 118. +doi: 10.1007/JHEP12(2019)118. arXiv: 1910.00538. +308 + +[249] +T. Weigand. Introduction to String Theory. 2012. url: http://www.thphys.uni- +heidelberg.de/~weigand/Skript-strings11-12/Strings.pdf. +[250] +S. Weinberg. The Quantum Theory of Fields, Volume 1: Foundations. Cambridge +University Press, May 2005. +[251] +S. Weinberg. The Quantum Theory of Fields, Volume 2: Modern Applications. Cam- +bridge University Press, May 2005. +[252] +P. West. Introduction to Strings and Branes. 1st edition. Cambridge University Press, +Aug. 2012. +[253] +E. Witten. ‘Non-Commutative Geometry and String Field Theory’. Nuclear Physics +B 268.2 (May 1986), pp. 253–294. doi: 10.1016/0550-3213(86)90155-0. +[254] +E. Witten. ‘A Note on the Antibracket Formalism’. Modern Physics Letters A 05.07 +(Mar. 1990), pp. 487–494. doi: 10.1142/S0217732390000561. +[255] +E. Witten. ‘On Background Independent Open-String Field Theory’. Physical Review +D 46.12 (Dec. 1992), pp. 5467–5473. doi: 10.1103/PhysRevD.46.5467. arXiv: hep- +th/9208027. +[256] +E. Witten. ‘Superstring Perturbation Theory Revisited’ (Sept. 2012). arXiv: 1209. +5461. +[257] +E. Witten. ‘The Feynman iϵ in String Theory’ (July 2013). arXiv: 1307.5124. +[258] +K. Wray. An Introduction to String Theory. May 2011. url: https://math.berkeley. +edu/~kwray/papers/string_theory.pdf. +[259] +X. Yin. ‘Aspects of Two-Dimensional Conformal Field Theories’. In: Proceedings of +Theoretical Advanced Study Institute Summer School 2017 "Physics at the Fun- +damental Frontier" — PoS(TASI2017). Vol. 305. SISSA Medialab, Sept. 2018, +p. 003. doi: 10.22323/1.305.0003. +[260] +J. Zinn-Justin. Quantum Field Theory and Critical Phenomena. 4th edition. Claren- +don Press, Aug. 2002. +[261] +B. Zwiebach. ‘Closed String Field Theory: An Introduction’ (May 1993). arXiv: hep- +th/9305026. +[262] +B. Zwiebach. ‘Closed String Field Theory: Quantum Action and the BV Master +Equation’. Nuclear Physics B 390.1 (Jan. 1993), pp. 33–152. doi: 10.1016/0550- +3213(93)90388-6. arXiv: hep-th/9206084. +[263] +B. Zwiebach. ‘Building String Field Theory around Non-Conformal Backgrounds’. +Nuclear Physics B 480.3 (Dec. 1996), pp. 541–572. doi: 10.1016/S0550-3213(96) +00502-0. arXiv: hep-th/9606153. +[264] +B. Zwiebach. ‘Oriented Open-Closed String Theory Revisited’. Annals of Physics +267.2 (Aug. 1998), pp. 193–248. doi: 10.1006/aphy.1998.5803. arXiv: hep-th/ +9705241. +[265] +B. Zwiebach. A First Course in String Theory. 2nd edition. Cambridge University +Press, Jan. 2009. +309 + +Index +#, 187, 190 +Σ0, see Riemann sphere +Σg, Σg,n, see Riemann surface +χg, χg,n, see Euler characteristics +| ↓⟩, see first-order system, vacuum +| ↑⟩, see first-order system, vacuum +⟨·⟩, 100 +ωg,n +p +, see Pg,n space, p-form +∼, 95 +†, see Hermitian adjoint +‡, see Euclidean adjoint +c, see conjugate state +t, see BPZ conjugation +[· · · ]g, see string product +{· · · }g, see fundamental vertex +1PI vertex, 222, 236 +region, 192 +1PR region, 192 +2d gravity, 51, 135 +adjoint, see Euclidean, Hermitian adjoint +AdS/CFT, 239 +Ag,n, see (off-shell) string amplitude +αn, see scalar field CFT, mode expansion +background independence, 238 +background metric, 32 +Batalin–Vilkovisky formalism, 23, 234, 288 +antibracket, 291 +BRST transformation, 291 +classical master equation, 291 +field redefinition, 292 +fields and antifields, 289 +gauge fixing, 292 +quantum BRST transformation, 293 +quantum master equation, 293 +bc CFT, see first-order CFT, reparametriza- +tion bc ghosts +Beltrami differential, 41 +Berkovits’ super-SFT, 256 +βγ CFT, see first-order CFT, superconformal +βγ ghosts +bosonic string CFT, 138 +L0, 138, 142 +complex parametrization, 141 +Hilbert space, 138 +level operator, 138, 142 +light-cone parametrization, 141 +boundary condition, 97 +Neveu–Schwarz (NS), 98 +Ramond (R), 98 +twisted, 98 +untwisted, 98 +BPZ conjugation, 97 +mode, 98 +state, 101 +BRST cohomology, see also string states, 69, +134 +absolute, 69, 139, 146 +relative, 70, 141, 144, 145, 147 +semi-relative, 70, 147 +two flat directions, 138–148 +BRST current, 135 +OPE, 135–136 +BRST operator, 135 +commutator, 137 +full +zero-mode decomposition, 147 +mode expansion, 136 +zero-mode decomposition, 136 +BRST quantization +change of gauge fixing condition, 287 +charge nilpotency, 287 +cohomology, 287 +Nakanishi–Lautrup auxiliary field, 286 +physical states, 287 +transformations, 286 +BRST SFT, see covariant SFT +BV formalism, see Batalin–Vilkovisky form- +alism +central charge, 33, 90, 95 +CFT, see conformal field theory +310 + +CISO, see conformal isometry group +CKV, see conformal Killing vector +classical solution +marginal deformation, 242 +closed string amplitude +Feynman diagram decomposition, 215 +on punctured moduli space, 178 +tree-level +3-point, 171 +4-point, 173, 200 +closed string fundamental vertex, 214 +g-loop +0-point, 221 +1-loop +0-point, 221 +1-point, 218 +properties, 223 +recursive construction, 217 +special vertices, 221 +tree-level +0-point, 221 +1-point, 221 +2-point, 221, 232 +3-point, 172 +4-point, 175 +closed string product, 223 +g = 0, n = 1, 230, 233 +ghost number, 223 +closed string states +dilaton Φ, 17 +Kalb–Ramond Bµν, 17 +level-matching, see level-matching +massless field, 17 +metric Gµν, 17 +tachyon T, 17 +physical, 148 +zero-mode decomposition, 176 +closed string vertex +fundamental identity, 223 +commutator +[Lm, Ln], 90 +[Lm, φn], 99 +CFT, 92–93 +complex coordinates, 72, 82 +cylinder, 83 +conformal +anomaly, see Weyl anomaly +dimension, 88 +factor, see Liouville field +spin, 88 +vacuum, see SL(2, C) vacuum +weight, 88 +conformal algebra, 79 +2d, see Witt algebra, Virasoro algebra +conformal field theory +classical, 84 +cylinder, 105 +definition, 88 +finite transformation, 88 +conformal isometry group, see also conformal +Killing vector, 78 +sphere, see SL(2, C) +conformal Killing +equation, 44, 79 +group volume, 46 +vector, 44, 62, 79 +conformal structure, 31 +conjugate state, 102 +conserved charge, 99, 268, 274 +CFT, 93 +conserved current, 268 +CFT, 93 +mode expansion, 99 +contracting homotopy operator, 139, 144 +contraction, 95 +conventions +complex coordinates, 268 +ghost number, 129 +spacetime momentum current, 109 +correlation function, 90 +quasi-primary operator, 90 +sphere 1-point (-), 91 +sphere 2-point (-), 91 +sphere 3-point (-), 91 +covariant SFT, see also classical (-), free (-), +quantum (-) +background independence, 239 +closed bosonic +1PI action, 236 +1PI gauge symmetry, 237 +classical action, 232 +classical equation of motion, 233 +classical gauge algebra, 233 +classical gauge transformation, 233 +fields and antifields, 235 +gauge fixed action, 230, 231 +inner product, 168 +normalization, 231 +quantum action, 235 +quantum BV master equation, 235 +open bosonic +inner product, 156 +parameters, 231 +renormalization, 231 +311 + +critical dimension, 14, 19, 48, 77, 135, 137 +superstring, 245 +cross-ratio, 91 +degeneration limit, 191 +diffeomorphism, 30 +group volume, 39, 44, 46 +infinitesimal (-), 30 +large (-), 31 +dual state, see conjugate state +ϵ (scalar action sign), 107 +ηξ ghosts, 246 +energy vacuum, 100 +energy–momentum tensor +finite transformation, 95 +mode expansion, 99 +Euclidean adjoint, 96–97 +mode, 98 +state, 101 +Euler characteristics, 30, 59 +extended complex plane ¯C, 81 +f◦, see CFT, finite transformation +factorization, see string amplitude +Faddeev–Popov +determinant, 44 +gauge fixing, see path integral +Fg,n, see propagator region +F1PR +g,n , see 1PR region +field space, see also path integral +DeWitt metric, 35 +inner-product, 34, 281 +norm, 34, 281 +field theory space, 239 +connection, 240 +Hilbert space bundle, 240 +first-order CFT, 119 +L0, 124, 128 +U(1) ghost current, 120, 129 +mode expansion, 124 +transformation law, 122 +U(1) symmetry, 119, 120 +action, 119 +boundary condition, 123 +BPZ conjugate +modes, 132 +vacuum, 132 +central charge, 121 +commutator, 125 +complex components, 119 +cylinder, 122 +energy normal ordering, 128 +energy–momentum tensor, 120 +mode expansion, 124 +equation of motion, 119 +Euclidean adjoint +modes, 132 +vacuum, 132 +Fock space, see Hilbert space, 131 +ghost charge, 121 +ghost number, 120, 129 +cylinder, 122 +Hilbert space +full, 131 +holomorphic, 131 +inner product, 133 +level operator, 124 +mode expansion, 123 +number operator, 124 +OPE, 121–123 +propagator, 120 +summary, 133 +vacuum +SL(2, C), 126 +conjugate, 133 +energy, 127–128 +Virasoro operators, 124, 129 +weight, 120 +zero-mode decomposition, 131 +zero-point energy, 127 +free covariant SFT +closed bosonic, 168 +action, 232 +classical action, 168 +equation of motion, 167 +gauge fixed action, 169, 228 +gauge fixed equation of motion, 159, +169, 231 +gauge transformation, 169 +open bosonic +BV action, 164 +classical action, 156, 157 +equation of motion, 155 +gauge fixed action, 159 +gauge transformation, 157, 165 +zero-mode decomposition, 156 +path integral, see string field path integ- +ral +free SFT, 140 +free super-SFT +action, 257, 258 +gauge transformation, 257, 258 +fundamental vertex, see also closed string (-) +312 + +interpretation, 231 +region, 191 +G, see spacetime ghost number +ˆgab, see background metric +gab, see worldsheet metric +ghost number, 52 +anomaly, 204, 250 +ghosts, see first-order CFT, reparametriza- +tion bc ghosts, superconformal βγ +ghosts +gluing compatibility, 190 +Green function, 59–61, 215 +tree-level 2-point, 221 +group +fundamental domain, 39 +volume, 38 +GSO symmetry, 247 +Hermitian adjoint, 96 +higher-genus Riemann surface +conformal group, 89 +Hilbert space +CFT, 100 +holomorphic factorization, 77, 251 +holomorphic/anti-holomorphic sectors, 83, 267 +I(z), I±(z), see inversion +iε-prescription, 262 +index +amplitude, 214 +Riemann surface, 194, 214 +inversion map, 87, 97 +ISO, see isometry group +isometry group, 79 +ker P1, see conformal Killing vector +ker P † +1 , see quadratic differential +Kg, see conformal Killing vector +Killing vector, 79 +L±, 99 +L∞ algebra, 234, 255 +left/right-moving sectors, 83, 267 +level-matching condition, 16, 70, 147, 168, +176, 184, 203, 205, 216, 228, 257 +ℓg,n, see string product +light-cone coordinates, 83 +Liouville +action, 47 +central charge, 48 +field, 32, 47 +free-field measure, 40 +theory, 88, 90, 264 +local coordinates, 23, 171, 179 +constraints, 184, 190 +global phase, 184, 202, 204 +global rescaling, 195 +reparametrization, 184, 202, 203 +transition function, 184, 185 +logarithmic CFT, 91 +mapping class group (MCG), see modular +group +marginal deformation +action, 239, 240 +correlation function, 240 +marked Riemann surface, see punctured Riemann +surface +metric +gauge +conformal (-), 32 +conformally flat (-), 32, 72 +flat (-), 32 +gauge decomposition, 39, 41, 43, 50 +gauge fixing, 31, 37 +uniformization gauge, 32 +local rescaling, see Weyl transformation +Mg, see moduli space +Möbius group, see SL(2, C) +mode expansion, 97 +Hermiticity, 98 +mode range, 97 +modular group, 31, 38 +moduli space, 21, 37 +complex coordinates, 76 +plumbing fixture decomposition, 190–194 +with punctures, 178 +momentum-space SFT +action, 261 +consistency, 264 +Feynman rules, 261 +finiteness +infinite number of states, 222, 261 +UV divergence, 222, 261 +Green function, 262 +interaction vertex, 261 +properties, 260 +string field expansion, 260 +Ngh, see ghost number +non-locality, 12, 260 +normal ordering, 103 +conformal (-), 103 +energy (-), 103 +313 + +mode relation, 105 +off-shell closed string amplitude, 179, 199 +contribution from subspace, 199 +tree-level +3-point, 172 +4-point, 218 +off-shell string amplitude, 23, 177 +conformal invariance, 172 +off-shell superstring amplitude, 250 +consistency, 250, 252 +factorization, 251–252 +PCO insertions, 250 +supermoduli space, 251 +old covariant quantization (OCQ), 146 +on-shell condition, 16, 60, 70, 92, 140, 144, +146, 147 +OPE, see operator product expansion +open string states +gauge field Aµ, 17 +gauge invariance, 17 +momentum expansion, 17 +tachyon T, 17 +operator product expansion, 94 +GG, 246 +T-primary, 95 +TG, 246 +TT, 95 +identity-primary, 94 +out-of-Siegel gauge constraint, 159 +p-brane, 11 +P1, 41, 42 +path integral +Faddeev–Popov gauge fixing, 36 +field redefinition, 283 +measure, 34, 282 +free-field, 34, 282 +ultralocality, 34, 49 +PCO, see picture changing operator +ˆPg,n, 184 +Pg,n space, 179 +p-form, 197, 198 +BRST identity, 203 +properties, 201–204 +0-form, 197 +1-form, 198 +coordinates, 185 +section, 179, 214 +vector, 185 +�Pg,m,n space, 250 +p-form, 250 +1-form, 250 +picture changing operator, 256 +picture number, 247 +anomaly, 247, 250 +plumbing fixture, 187, 208, 251 +p-form, 209 +ghost 1-form, 209 +moduli, 187 +non-separating, 190, 212 +separating, 187–190, 208, 222 +vector field, 209 +Polyakov path integral, 29 +complex representation, 75 +Faddeev–Popov gauge fixing, 36, 61 +primary operator/state, 88 +finite transformation, 88 +weight-0, 91 +projector +on-shell, ker L0, 140 +propagator, 139 +closed bosonic, 212, 215–217 +with stub, 222 +closed string, 174 +NS sector, 251 +open bosonic, 158 +R sector, 252 +Schwinger parametrization, 22, 216 +superstring, 256 +propagator region, 191 +puncture, see Riemann surface +punctured Riemann surface, 178 +Euler characteristics, 178 +parametrization, 182–183 +quadratic differential, 42 +quasi-primary operator/state, 88 +r(Σg,n), see index +radial quantization, 92 +reparametrization bc ghost +zero-mode, 69 +reparametrization bc ghosts, see also first- +order CFT, 51 +action, 51, 76 +central charge, 47 +equation of motion, 51 +zero-mode, 54 +Rg,n, see off-shell string amplitude contribu- +tion +Riemann sphere S2, 81 +complex plane map, 81 +cylinder map, 83 +314 + +Riemann surface +degeneration, 23 +g = 0, see Riemann sphere +genus, 29 +puncture, 20, 58 +scalar field CFT +L0, 115–116 +U(1) current, 108, 109 +U(1) symmetry, 108 +action, 107, 109 +boundary condition +periodic, 115 +BPZ conjugate +modes, 118 +vacuum, 118 +central charge, 111 +commutator, 116 +complex components, 109 +complex plane, 109 +cylinder, 110 +dual position, 114 +energy–momentum tensor, 107, 110 +equation of motion, 107, 109 +Euclidean adjoint +modes, 118 +vacuum, 118 +Fock space, 117 +Hilbert space, 117 +inner product, 118 +level operator, 115 +mode expansion, 16, 113–115 +momentum, 108, 110, 114, 115 +normal ordering, 113 +number operator, 115 +OPE, 111–113 +periodic boundary condition, 107 +position (center of mass), 114 +propagator, 107 +topological current, 108, 110 +vacuum, 117 +conjugate, 118 +vertex operator, 110 +Virasoro operators, 115 +winding number, 109, 110, 114, 115 +zero-mode, 114 +scattering amplitude, 59–61 +tree-level 2-point (-), 60 +Schwarzian derivative, 95 +Schwinger parameter, see propagator +section of Pg,n +generalized, 224 +overlap, 224 +SFT, see string field theory +Siegel gauge, 141, 158, 165, 169 +SL(2, C) group, 86 +SL(2, C) vacuum, 100 +S-matrix, see scattering amplitude +spacetime ghost number +closed string, 227 +open string, 161 +spacetime level-truncated action +open bosonic, 165 +gauge transformation, 166 +spacetime momentum, see also scalar field +CFT, 109 +spurious pole, 250, 252–255 +state (CFT) +bra, 101 +ket, 100 +state–operator correspondence, 100 +Stokes’ theorem, 75 +string +gauge group, 19 +interactions, 12, 20 +orientation, 19 +parameters, 21 +properties, 14 +string amplitude, 20 +CKV gauge fixing, 62 +closed string, see closed string amplitude +conformal invariance, 171 +divergence, 22 +factorization, 174, 208, 214 +ghost number, 204 +g-loop vacuum (-), 34, 49, 52, 54, 55, 76 +matching QFT, 59–60, 64 +normalization, 55, 58 +properties, 204–207 +pure gauge states decoupling, 206 +section independence, 206 +tree-level 2-point (-), 64–67 +string amplitudegn +g-loop n-point (-), 59, 63, 64 +string coupling constant, 21 +string Feynman diagram, see also momentum- +space SFT, 172 +1PR diagram, 211, 222 +change of stub parameter, 222 +Feynman rules, 213 +intermediate states +ghost number, 212, 213 +momentum, 212, 213 +IR divergence, 231 +315 + +loop diagram, 213 +string field, see also string states +closed bosonic, 168 +classical, 232 +expansion, 226, 235 +quantum, 235 +expansion, 152 +functional, 150 +gauge fixing, see Siegel gauge +ket representation, 151 +momentum expansion, 25, 151, 260 +open bosonic, 155 +classical, 156 +expansion, 160 +Nakanishi–Lautrup auxiliary field, 165 +parity, 156 +quantum, 161 +reality condition, 156 +position representation, 151 +string field path integral +Faddeev–Popov gauge fixing, 162 +free covariant open bosonic string, 162 +string field theory, 25 +construction, 23 +string states, see also closed, open (-), 14, +175–177 +dual, 176 +off-shell, 175 +on-shell, see on-shell condition +physical, see BRST cohomology +resolution of identity, 176 +string theory +background independence, 238 +consistency, 64, 264 +motivations, 11–13 +structure constant (CFT), 91 +stub, 194–195 +Feynman diagram, 222 +parameter, 194 +stub parameter, 232, 261 +super-SFT, 255 +superconformal βγ ghosts, see also first-order +CFT, 246 +bosonization, 246 +conformal weights, 246 +energy–momentum tensor, 246 +OPE, 246 +superstring, 18, 245 +BRST current, 247 +heterotic (-), 19, 245 +Hilbert space, 248 +large, 248 +picture number, 248 +small, 248 +motivations, 18 +normalization, 248 +type I (-), 19 +type II (-), 19 +superstring field +auxiliary field, 257 +constrained Ramond field, 256 +large Hilbert space, 258 +Ramond field, 255 +small Hilbert space, 256 +supersymmetry, 18 +surface state, 200, 209 +T-duality, 114 +tachyon, 17 +instability, 18 +Teichmüller deformation, 39, 41, 42 +Teichmüller space, 37 +ultralocality, see path integral measure +Verma module, 102 +vertex operator, see also scalar field CFT, see +also string states +integrated, 58 +unintegrated, 64 +Weyl invariance, 61 +vertex state, 223 +vertical integration, 253 +Vg,n, see fundamental vertex +V1PI +g,n , see 1PI vertex +Virasoro algebra, 90 +Virasoro operators, 90, 99 +Hermiticity, 99 +Weyl +anomaly, 35, 47, 49 +ghost, 52, 68 +group volume, 39 +symmetry, 31, 52 +Wick rotation, 262, 266 +generalized, 263 +Wick theorem, 103 +winding number, see also scalar field CFT +Witt algebra, 85 +worldsheet +action +Einstein–Hilbert (-), 56 +gauge-fixing (-), 68 +Nambu–Goto (-), 29 +316 + +Polyakov (-), 29 +sigma model (-), 30 +boundary conditions, 14 +CFT, 14, 18 +classification, 14 +cosmological constant, 48 +cylinder, 83 +energy–momentum tensor, 33 +trace, 33 +ghosts, see reparametrization ghosts +Nakanishi–Lautrup auxiliary field, 68 +path integral, see Polyakov path integral +Riemann surface, 20, 29 +symmetry, 30 +background diffeomorphisms, 50 +background Weyl, 50 +BRST, 68 +diffeomorphisms, 30 +Weyl, 30 +worldsheet metric, 29 +worldvolume description, 12 +X, see picture changing operator +Zamolodchikov metric, 91 +zero-mode, 269, 283 +zero-point energy, 101 +317 +