J H E P 1 0 ( 2 0 1 8 ) 1 7 9 Published for SISSA by Springer Received: August 22, 2018 Accepted: October 18, 2018 Published: October 29, 2018 Finite deformations from a heterotic superpotential: holomorphic Chern-Simons and an L1 algebra Anthony Ashmore,a Xenia de la Ossa,a Ruben Minasian,b Charles Strickland-Constablec;d and Eirik Eik Svanese;f aMathematical Institute, University of Oxford, Andrew Wiles Building, Radcli e Observatory Quarter, Woodstock Road, Oxford, OX2 6GG, U.K. bInstitut de Physique Theorique, Universite Paris Saclay, CEA, CNRS, F-91191, Gif-sur-Yvette, France cSchool of Mathematics, University of Edinburgh, James Clerk Maxwell Building, Peter Guthrie Tait Road, Edinburgh, EH9 3FD, U.K. dSchool of Physics, Astronomy and Mathematics, University of Hertfordshire, College Lane, Hat eld, AL10 9AB, U.K. eDepartment of Physics, King's College London, London, WC2R 2LS, U.K. fThe Abdus Salam International Centre for Theoretical Physics, 34151 Trieste, Italy E-mail: ashmore@maths.ox.ac.uk, delaossa@maths.ox.ac.uk, ruben.minasian@ipht.fr, c.strickland-constable@herts.ac.uk, eirik.svanes@kcl.ac.uk Abstract: We consider nite deformations of the Hull-Strominger system. Starting from the heterotic superpotential, we identify complex coordinates on the o -shell parameter space. Expanding the superpotential around a supersymmetric vacuum leads to a third- order Maurer-Cartan equation that controls the moduli. The resulting complex e ective action generalises that of both Kodaira-Spencer and holomorphic Chern-Simons theory. The supersymmetric locus of this action is described by an L3 algebra. Keywords: Superstring Vacua, Flux compacti cations, Superstrings and Heterotic Strings, Di erential and Algebraic Geometry ArXiv ePrint: 1806.08367 Open Access, c The Authors. Article funded by SCOAP3. https://doi.org/10.1007/JHEP10(2018)179 J H E P 1 0 ( 2 0 1 8 ) 1 7 9 Contents 1 Introduction 1 2 The Hull-Strominger system and a heterotic superpotential 4 2.1 N = 1 heterotic vacua and the Hull-Strominger system 4 2.2 The Atiyah algebroid and a holomorphic structure 5 3 The o -shell parameter space 7 3.1 F -term conditions from the superpotential 9 3.2 Constraints from the SU(3) structure and holomorphicity 10 4 Higher-order deformations 12 4.1 The superpotential 12 4.2 A Maurer-Cartan equation from the holomorphic structure 13 4.3 Including the bundle 14 4.4 Vanishing of the superpotential 16 5 Moduli and an L3 algebra 17 5.1 Quasi-isomorphism to a natural holomorphic L3 algebra 18 5.2 An L1 eld equation 19 6 A reduced L3 algebra and an e ective action 20 6.1 Integrating out b 20 6.2 E ective eld theory and Yukawa couplings 22 7 Conclusions 24 A Conventions 26 A.1 Heterotic supergravity 27 A.2 Holomorphic structure 27 A.3 L1 structure 28 B Comments on heterotic ux quantisation 30 B.1 A toy example: the abelian bundle 30 B.2 The two-form gerbe example 31 B.3 Heterotic ux quantisation 32 B.4 Deforming the system and a well-de ned global two-form 34 C The o -shell N = 1 parameter space and holomorphicity of 35 C.1 SU(3) SO(6) structures in the NS-NS sector 36 C.2 SU(3) SO(6 + n) structures in heterotic supergravity 39 C.3 The o -shell hermitian structure on V 40 { i { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 D Comments on D-terms 42 D.1 Massless deformations 42 D.2 Including bundle moduli 45 D.3 Polystable bundles 48 D.4 Full Maurer-Cartan equations 49 E Massless moduli 50 1 Introduction A full understanding of the parameter space of string theory is an outstanding mathemat- ical challenge and would lead to powerful constraints on the landscape of string models. Of the various limits of string theory, the heterotic string has been the focus of much phenomenology thanks to the relative ease with which one can engineer four-dimensional theories with chiral fermions and the Standard Model gauge group [1{8]. Much of this work has been on models where the internal manifold is Calabi-Yau, mostly because such spaces can be constructed using algebraic geometry and then used for compacti cations without knowledge of their explicit metrics. Calabi-Yau compacti cations are not the most general way to obtain an N = 1 theory in four dimensions that admits a Minkowski vacuum. The general solution to O( 0) is given by compactifying on a complex three-fold X with H ux and a gauge bundle that satisfy an anomaly cancellation condition. The conditions on the geometry and uxes for such a solution are known as the Hull-Strominger system [9, 10]. Known solutions to this system include Calabi-Yau spaces with bundles and a small number of honestly non-Kahler geometries. Generically, a given solution of the Hull-Strominger system will admit deformations of the geometry, ux and bundle that remain N = 1 solutions | these deformations are known as moduli. These moduli appear in the massless spectrum of the low-energy theory, so it is important that we understand the moduli space of a given compacti cation. The moduli spaces of Calabi-Yau compacti cations at zeroth order in 0 are well un- derstood using the language of special geometry. Until recently the general case had not been tackled | this might come as a surprise. Certainly in type II theories the condi- tions for an N = 1 Minkowski solution are suciently complicated (thanks to branes and other ingredients) that their moduli spaces might not admit a general formulation. In the heterotic case, the underlying geometry is relatively straightforward. One might have expected that the gauge sector and anomaly conditions complicate matters somewhat, but that the moduli space might still be understood. Starting with [5, 11{15], this gap is now being lled. (See also [16{19] for worldsheet approaches.) In nitesimally, the moduli space is characterised by the existence of a holomorphic structure D on a bundle Q over the three-fold X. The fact that the moduli space is nite dimensional is intimately connected to this holomorphic structure and the Bianchi identity { 1 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 for the ux. The in nitesimal moduli are captured by the cohomology H (0;1) D (Q), where Q is de ned by a series of extensions. In this way, the complex structure, hermitian and bundle moduli are combined in a single structure. Furthermore, one can de ne the analogue of special geometry for these general heterotic compacti cations and nd the metric on the moduli space [20, 21]. A natural question to ask is whether one can understand the moduli spaces to higher order. If we think about deformations of a complex structure, we know the in nitesimal moduli are given by H (0;1) @ (T (1;0)X), while the higher-order deformations satisfy a Maurer- Cartan equation. A similar thing happens for bundle deformations [22], and simultaneously deformations of the bundle and complex structure [23, 24]. In this way, moduli can be obstructed at higher orders and can give non-zero contributions to the superpotential of the four-dimensional e ective theory. The aim of this work is to derive the corresponding conditions on the moduli for the Hull-Strominger system at higher orders. In other words, we want to derive the conditions on the moduli when they describe a small but nite deformation of the original heterotic solution. There are a number of ways one might go about this. One path would be to start with the equations of the Hull-Strominger system and deform the various elds. The deformed elds should still satisfy the Hull-Strominger system (as it describes the most general solution) so one can rewrite the system of equations as conditions on the deformations themselves. This is similar to the path taken in [20, 21]. Our approach will be compli- mentary. It has been shown that supersymmetry of the heterotic system can be described using a four-dimensional superpotential [21, 25, 26]. The vanishing of the superpotential and its rst derivative imposes the F -term conditions in the four-dimensional theory and leads to an N = 1 Minkowski vacuum. Our plan is to deform the elds that appear in the superpotential and then read o the conditions on the moduli for the superpotential and its derivative to vanish. These two approaches will be shown to be equivalent in a future publication [27]. A particular advantage of proceeding this way is that one can use the knowledge that the superpotential is a holomorphic function of the moduli elds to streamline the problem. In addition to the usual N = 1 lore that the superpotential is holomorphic, we give an argument that the superpotential is holomorphic on the space of moduli elds without requiring that they give a solution to the Hull-Strominger system. This is equivalent to saying that the o -shell parameter space | the space of SU(3) structures, B elds and gauge bundles | is a complex space and the superpotential is a holomorphic function of these parameters. We outline how this follows from generalised geometry where N = 1 NS-NS compacti cations are described by a generalised SU(3)  SU(4) structure [28]. We identify the invariant (holomorphic) object which characterises this structure and nd that the complex coordinates on the space of structures match with the usual complex structure, complexi ed hermitian and bundle deformations. We show that the conditions on the moduli elds from the vanishing of the superpoten- tial and its rst derivative can be written as a pair of third-order Maurer-Cartan equations using the holomorphic structure and a number of brackets. Moreover, we show that the { 2 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 superpotential itself can be rewritten using these operators in a Chern-Simons-like form: W = Z X hy;Dy 1 3 [y; y] @bi ^ ; (1.1) where y describes the complex, hermitian and bundle moduli, and b is a (0; 2)-form. W is written using a holomorphic structure D, a pairing h; i on the moduli elds, and a bracket [; ]. The brackets can be understood as coming from an underlying holomorphic Courant algebroid that describes the combined deformations of the complex structure, metric and uxes [29, 30]. We then show that the supersymmetry conditions can be recast in terms of an L3 algebra. We outline how the L3 algebra gives a C1 resolution of the underlying holo- morphic Courant algebroid. The natural L3 eld equation reproduces the supersymmetry conditions, and the L3 structure gives the gauge symmetries of the moduli space in a compact form. It is known that generic deformation problems have a description in terms of L1 structures, so it is not unexpected that our moduli elds are governed by one. What is unexpected is that the structure truncates at nite order leaving us with an L3 algebra. Why does the deformation truncate in our case? A generic deformation problem can be parametrised in many equivalent ways | some may truncate at nite order while oth- ers do not. Essentially, the structure of the heterotic system and its formulation using a superpotential guides us to pick a \nice" parametrisation. Said another way, we know from supergravity that the superpotential should be a holomorphic function of the param- eters. Thus when we express the superpotential in the obvious complex coordinates on the parameter space, we get the most natural way to package the deformation problem. We begin in section 2 with a review of the Hull-Strominger system and the description of its in nitesimal moduli in terms of a holomorphic structure as in [14]. In section 3 we discuss the o -shell parameter space of the theory and give the complex coordinates on the parameter space. We show how the F -term conditions follow from a heterotic superpotential to set the scene for the higher-order deformations. In section 4 we examine the higher-order deformation problem and nd the system of equations that govern the moduli of the Hull-Strominger system. We show how this can be written in terms of the holomorphic structure D and a bracket [; ] arising from a holomorphic Courant algebroid. In section 5 we rewrite the equations that govern the moduli in terms of an L3 structure. We give the various multilinear products `k that de ne the L3 structure and discuss how various properties, such as the moduli equations and gauge symmetries, are naturally encoded in this L3 language. In section 6 we discuss how the system simpli es under various assumptions and comment on how the e ective eld theory is encoded in our language. We nish with a discussion of some open questions and avenues for future work. In the appendices, we lay out our conventions, include a few comments on how ux quantisation works in the heterotic theory, discuss the o -shell parameter space in terms of generalised geometry, show that the D-term conditions do not a ect the moduli problem and review how the massless moduli are captured by the the cohomology of the holomor- phic structure. { 3 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 2 The Hull-Strominger system and a heterotic superpotential We begin with a review of the Hull-Strominger system [9, 10] and the description of its in nitesimal moduli using a holomorphic structure [14, 15, 31]. 2.1 N = 1 heterotic vacua and the Hull-Strominger system The Hull-Strominger system is a set of equations whose solutions are supersymmetric Minkowski vacua of heterotic string theory to order O( 0). The ten-dimensional solution is a product of four-dimensional Minkowski space with a six-dimensional complex manifold X. X admits a vector bundle V with connection A whose curvature F is valued in EndV . The tangent bundle TX of X also admits a connection  whose curvature R is valued in EndTX. X admits an SU(3) structure de ned by a nowhere vanishing spinor  or, equivalently, a non-degenerate two-form ! and a nowhere vanishing three-form that are compatible ! ^ = 0; ik k2 ^ = 1 3! ! ^ ! ^ !: (2.1) The invariant objects are de ned by bilinears of the spinor as !mn = i y mn; mnp = T mnp; (2.2) where we are free to normalise the spinor so that k k2 = 8. In what follows it will be useful to de ne a three-form which is related to by a dilaton factor as = e2 : (2.3) Supersymmetry of the vacuum follows from the vanishing of the supersymmetry vari- ations of the fermionic elds, given in equations (A.7){(A.9). To rst order in 0, these conditions are equivalent to the Hull-Strominger system: d = 0; (2.4) i(@ @)! = H := dB + 0 4 (!CS(A) !CS()); (2.5) ^ F = 0; (2.6) !yF = 0; (2.7) d(e2! ^ !) = 0; (2.8) where !CS is the Chern-Simons three-form for the connection, !CS(A) = tr  A ^ dA+ 2 3 A ^A ^A  : (2.9) The closure of from (2.4) implies that the manifold is complex with a holomorphically trivial canonical bundle, while condition (2.8) tells us that X is conformally balanced. { 4 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 Conditions (2.6) and (2.7) mean V is a polystable holomorphic bundle1 so the curvature satis es the hermitian Yang-Mills equations. Finally, (2.5) de nes the ux H in terms of the heterotic B eld and the anomaly cancellation condition, and links it with the intrinsic torsion of the SU(3) structure. The corresponding Bianchi identity is2 dH = 0 4 (trF ^ F trR ^R): (2.10) This set of conditions de nes what one might call a heterotic SU(3) structure. Upon considering the four-dimensional N = 1 theory that would follow from compact- ifying on such a solution, the Hull-Strominger system naturally splits into F - and D-term conditions. As discussed in [26], the F -term equations are d = 0; i(@ @)! = H; ^ F = 0: (2.11) It is these equations that the heterotic superpotential reproduces. The remaining equations of the Hull-Strominger system are the conformally balanced condition and the Yang-Mills equations, referred to as the D-term equations. Modulo certain mild assumptions on the geometry, the in nitesimal deformations are parametrised by the cohomology H (0;1) D (Q), where D and Q are to be de ned below. This cohomology is reviewed in appendix E. Under in nitesimal deformations, the D-term equa- tions x a representative of a certain cohomology class [14], and so should be thought of as gauge xing conditions that do not a ect the moduli problem. This is of course expected from the four-dimensional N = 1 supergravity point of view [33{35]. In appendix D we show that preserving the D-term conditions for nite deformations also amounts to xing a gauge. One might worry about Fayet-Iliopoulos terms appearing, but these are in fact accounted for by modding out by D-exact terms, as shown in [14]. 2.2 The Atiyah algebroid and a holomorphic structure The vector bundle V is hermitian in agreement with (0; 2) supersymmetry on the world- sheet [36]. The curvature F of the bundle is given by F = dA+A ^A; (2.12) where A is a one-form connection valued in End V . The exterior derivative on V twisted by A is dA := d + [A; ]; (2.13) 1More precisely, it is the complex vector bundle VC (de ned in appendix C.3) that is a holomorphic bundle. 2The curvature R in the Bianchi identity is the curvature of a connection on TX, satisfying its own hermitian Yang-Mills conditions in order for the equations of motion to be ful lled [32]. To O( 0), this connection is r, given by taking the connection in (A.7) with the opposite sign for H. { 5 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 where the action of the bracket on a p-form is [A; ] = A ^ (1)p ^A: (2.14) A holomorphic structure on V is xed by the (0; 1) component of dA, which we denote @A. This operator squares to zero if the bundle is holomorphic, that is F(0;2) = 0. Moreover, the Bianchi identity for the curvature is simply @AF = 0: (2.15) A deformation of the Hull-Strominger system corresponds to simultaneous deforma- tions of the complex structure, hermitian structure and gauge bundle. Taking each of these in isolation is not sucient. In particular, deformations of the hermitian structure alone lead to an in nite-dimensional moduli space. It is surprising that if one considers the full deformation problem together with the anomaly cancellation condition, one nds a nite-dimensional moduli space. Of course, this is what one would expect from string theory, but the precise way in which this happens is rather remarkable. As discussed in [14], the in nitesimal moduli of the Hull-Strominger system are cap- tured by deformations of a holomorphic structure. The holomorphic structure D acts on a bundle Q. Locally Q is given by Q ' T (1;0)(X) EndV  EndTX  T (1;0)X: (2.16) Globally, Q is de ned by an extension3 0! T (1;0)(X)! Q! Q1 ! 0; (2.17) where the bundle Q1 is de ned by 0! EndV  EndTX ! Q1 ! T (1;0)(X)! 0: (2.18) The holomorphic structure D on Q is a derivative4 D : (0;p)(Q)! (0;p+1)(Q); (2.19) where D 2 = 0 if and only if the Bianchi identities for H, F and R are satis ed. The Hull- Strominger system is then equivalent to the data of the extension bundle Q, the nilpotent holomorphic structure D, polystability of V and TX and the conformally balanced condi- tion on X. The in nitesimal deformations of the holomorphic structure are simply elements of the D-cohomology of Q-valued (0; 1)-forms | H(0;1) D (Q). As shown in [14], this is also the moduli space of heterotic SU(3) structures. We give a short review of this in appendix E. For the rest of the paper, we make a eld rede nition to absorb the explicit 0 dependence B ! 0 4 B; ! ! 0 4 !: (2.20) 3Full details can be found in [14]. 4A similar operator has appeared in the context of generalised Kahler geometry [29]. { 6 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 One can restore the proper factors of 0 by the inverse transformations. We will also suppress the connection on TX | we can reintroduce it in what follows by treating TX as part of the gauge bundle and de ning the bundle metric on the TX subspace to be negative de nite so that the Bianchi identity comes with a negative sign for the tr R ^R term. The main aim of this work is to understand what happens for nite deformations. In particular we will see the holomorphic structure is an important ingredient in describing higher-order deformations. First let us discuss the o -shell parameter space and how the heterotic SU(3) structure can be rephrased using a superpotential. 3 The o -shell parameter space We now show that a subset of the system corresponding to F -term conditions can be de- rived from a superpotential. We then discuss how deformations of the geometry, ux and bundle can be parametrised using the observation that the superpotential is a holomor- phic function. The four-dimensional e ective theory that one nds after compactifying the heterotic string on an SU(3) structure manifold is controlled by a superpotential [21, 25, 26]. The superpotential W is given in terms of the ux H and the SU(3) invariant forms by5 W = Z X (H + i d!) ^ : (3.1) As we will review, given the SU(3) structure relations (2.1) and the de nition of H in (2.5), W = W = 0 reproduces the F -term conditions of the Hull-Strominger system [26]. Notice that W = 0 requires us to vary the superpotential over some space of eld con gurations. We need to understand what this space is in order to nd how the superpo- tential behaves when we perform a nite deformation of the background elds. We pause brie y to distinguish between this parameter space and the moduli space of solutions to the Hull-Strominger system. The parameter space or space of eld con gurations Z is the space of SU(3) structures, B elds and hermitian gauge bundles on the real manifold X. The SU(3) structure is equivalent to the existence of a nowhere-vanishing spinor so that on this space of eld con gurations the heterotic theory admits an \o -shell" N = 1 supersymmetry of the kind discussed in [37, 38]. These elds do not necessarily solve the Hull-Strominger system and so we often refer to them as o -shell eld con gurations. This is the space over which the superpotential is varied. The moduli space M of the Hull-Strominger system is a subspace of Z on which the elds also solve the Hull-Strominger system. This set of elds is what one usually means by moduli, and we will often refer to them as on-shell con gurations. Another way of saying this is that the superpotential and its derivatives vanish when evaluated on M. As W is a superpotential for a four-dimensional N = 1 theory we expect it to be a holomorphic function. The holomorphicity of W is a powerful tool for understanding deformations of the Hull-Strominger system and we will see later how its presence greatly 5Here we have scaled away an overall factor of 0=4 that comes from the eld rede nition in (2.20). { 7 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 simpli es the problem. On the supersymmetric locus (on-shell), it is known that the superpotential is a holomorphic function of the moduli elds [21] | this is simply the statement that anti-holomorphic derivatives of the superpotential vanish on imposing the F -term conditions. Physics goes further than this and insists that W is a holomorphic function of the o -shell eld con gurations | the o -shell eld space Z must admit complex coordinates and W must be a holomorphic function of these coordinates. In other words, the three-form and the particular combinations of B + i! and A that appear in (3.1) must be parametrised by these complex coordinates. We outline a proof that such complex coordinates exist and that is holomorphic on the parameter space Z in appendix C using the formalism of generalised geometry. We also discuss how the hermitian structure on V survives o -shell. For completeness, one should really show that W itself can be expressed as a holomorphic function of the object ~ that we de ne in appendix C | we leave this for a future work. Note that on-shell, is also holomorphic as a function of the complex coordinates of X. When we talk of being holomorphic we are instead referring to its dependence on the coordinates of the o -shell parameter space. Let t and t denote holomorphic and anti-holomorphic coordinates on the parameter space Z. The corresponding holomorphic and anti-holomorphic variations are  = 1X n=1 1 n! tnDnt ;  = 1X n=1 1 n! tnDnt ; (3.2) where D is a covariant derivative on the parameter space [20]. As we discuss in appendix C.3, o -shell the gauge bundle admits a real hermitian connection valued in V . This decomposes into (1; 0)- and (0; 1)-forms, with a corresponding decomposition of the Chern-Simons form. Not all components of the Chern-Simons form appear in W (as it is wedged with ); only the (0; 1)-form components of A contribute. It is this component of the connection that is the complex coordinate on the o -shell parameter space. A holomorphic deformation of the connection is then given by a (0; 1)-form valued in EndV , which we denote : A 7! A+ A = A+ : (3.3) We show in appendix C.3 that one can write the Chern-Simons form using connections valued in V , VC, V C or a combination of these. From this it is clear that it is equivalent to work with (0; 1)-forms valued in End V or with (0; 1)-forms valued in End VC. A deformation of the complex structure is parametrised by a (1; 0)-vector valued (0; 1)- form,  2 (0;1)(T (1;0)X), also known as a Beltrami di erential. The complex coordinates on X that de ne the complex structure deform to dza ! dza + dza = dza + a: (3.4) In nitesimally, the deformed complex structure is J + 2i. For a small but nite deforma- tion, the holomorphic three-form becomes [39, 40] ! +  = + { + 1 2 {{ + 1 3! {{{ ; (3.5) { 8 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 where the variations of in coordinates are { = 1 2 abc a ^ dzbc; {{ = abc a ^ b ^ dzc; {{{ = abc a ^ b ^ c: (3.6) The fact that a variation of is completely captured by  without needing  is an indication that is a holomorphic function of the coordinates of the o -shell parameter space. Note that the variation of can in principle have a (3; 0) component. However, the (3; 0) part should be interpreted as a Kahler transformation and so is not part of the physical moduli. Another way of saying this is that  in (3.2) is built from covariant derivatives on the parameter space [20]. Restricted to variations of the complex structure, the (3; 0) component is attributed to a connection in the usual way. The holomorphic deformations of the hermitian and B eld moduli are (B + i !)(1;1) and B(0;2); (3.7) where a subscript (p; q) denotes the type with respect to undeformed complex structure. Here B is a combination of variations of the B eld and the exact term in the variation of the Chern-Simons term, given by B = B + tr A ^A; (3.8) up to a d-closed two-form. The Green-Schwarz mechanism ensures B is gauge invariant. As we show in appendix B, ux quantisation then implies that B is a globally de ned two-form, so it can indeed be a modulus. As we will see, the (0; 2) component of ! is actually xed in terms of the other moduli by the SU(3) relations, but it will be convenient to package this with B(0;2) into (B + i !)(0;2). 3.1 F -term conditions from the superpotential Let us review how one derives the F -term conditions from the superpotential. We take  to be an in nitesimal deformation of the elds, leading to a corresponding variation of the superpotential W W = Z X 2 tr A ^ F + d(B + i !) ^ + Z X (H + i d!) ^  : (3.9) The F -term conditions come from requiring that both W and W vanish for generic values of the moduli. For arbitrary B + i ! and A, the vanishing of W requires @ = d = 0; F ^ = 0: (3.10) With this in mind, the vanishing of W for arbitrary  of type (2; 1) impliesH(1;2) = i @!. Using the previous conditions, the vanishing of W itself reduces to H(0;3) = @ for some { 9 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 (0; 2)-form . The Bianchi identity for H then implies H(0;3) = 0, giving us the nal F -term condition: H = i(@ @)!: (3.11) With this we see W = W = 0 reproduces the F -term equations of the Hull- Strominger system. Our plan is to extend this discussion to understand nite deformations around a su- persymmetric solution. First we need to understand how the requirement of an SU(3) structure and holomorphicity of the superpotential constrain the possible deformations. 3.2 Constraints from the SU(3) structure and holomorphicity The existence of an SU(3) structure is part of the data that goes into the superpotential, so the deformed geometry should also de ne an SU(3) structure. Another way of saying this is that the o -shell parameter space on which the superpotential is varied is the space of SU(3) structures (plus bundles, and so on). This means the SU(3) structure compatibility condition must still hold: (! + !) ^ ( +  ) = 0; (3.12) where  is a nite holomorphic variation. Expanding this out according to complex type, we nd 0 = ! + (!)(2;0) + (!)(1;1) + (!)(0;2)  ^  + { + 1 2 {{ + 1 3! {{{  : (3.13) Upon contracting this with , we see this equation xes the (0; 2) component of ! in terms of the other deformations: (!)(0;2) = {! + {(!)(1;1) 1 2 {{(!)(2;0): (3.14) Now consider an anti-holomorphic variation  under which does not vary as it is a holomorphic on the parameter space. We then note that as an anti-holomorphic deformation does not change the complex structure (as  = 0) the SU(3) compatibility condition reduces to ! ^ = 0: (3.15) From this we see (!)(0;2) = 0 and so, taking a conjugate, (!)(2;0) = 0. Combined with the previous result of a holomorphic variation of the compatibility condition, we have (!)(2;0) = 0; (!)(0;2) = {! + {(!)(1;1): (3.16) From this we see the (0; 2) component of the variation of ! is xed by the complex structure and hermitian moduli. We can also play the same trick with the superpotential itself. The superpotential is a holomorphic function of the moduli so an arbitrary anti-holomorphic variation of it must vanish without having to impose the supersymmetry conditions | it must vanish { 10 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 \o shell". Let us turn o the gauge sector for now, and consider an anti-holomorphic variation of W : W = Z X d(B + i !) ^ = Z X (B + i !) ^ d : (3.17) As W is a holomorphic functional, this must vanish for all anti-holomorphic deformations without needing to impose the F -term conditions. Without imposing integrability of the complex structure, generically we have d 2 (3;1)(X) (2;2)(X). For an anti-holomorphic variation of the superpotential to vanish, it is sucient that (B + i !)(1;1) = 0; (B)(0;2) = 0; (3.18) where we have used the rst condition from (3.16) to remove (!)(0;2). This agrees with (3.7) where we stated that the holomorphic combinations are the (1; 1) and (0; 2) components of B + i! (see also discussion in appendix C.1). Taking a conjugate of these conditions we have (B i !)(1;1) = 0; (B)(2;0) = 0: (3.19) Taken together these give (B)(1;1) = i (!)(1;1); (B)(2;0) = (!)(2;0) = 0: (3.20) For what follows, it is useful to de ne ~b = (B + i !)(0;2); x = i (!)(1;1) = (B)(1;1); (3.21) which are our complex coordinates on the parameter space. One can repeat this exercise with gauge sector turned on. W is a holomorphic function of the moduli so it does not change for arbitrary anti-holomorphic variations A. For W = 0 to hold at a generic o -shell point in eld space, we nd it is sucient that (A)(0;1) = 0. This implies (A)(1;0) = 0, so the holomorphic deformations correspond to A = 2 (0;1)(EndV ): (3.22) In other words, the holomorphic coordinate on the parameter space is , in agreement with (3.3). Furthermore, one sees that the holomorphic deformations of the complexi ed hermitian moduli are the (1,1) and (0,2) components of B + i!. As an aside, we note that there is a schematic way to see that is a holomorphic func- tion of the parameter space coordinates. Consider a generic anti-holomorphic deformation of the superpotential around a point in moduli space where the holomorphic top-form is closed, d = 0: W = Z X (H + i d!) ^ + Z X d(B + i !) ^ : (3.23) For an in nitesimal deformation, the second term can be dropped, and a sucient condition for the rst term to vanish for generic H and ! is that  = 0; (3.24) { 11 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 in nitesimally. Note that this also kills the second term in (3.23) at second order in perturbation theory. A sucient condition for W to vanish at this order is hence that  = 0 to this level as well. This argument can be continued ad in nitum, and we are left with condition (3.24), at least for nite deformations away from a closed . We will assume that this condition holds true at generic o -shell points in the parameter space. Stronger evidence for this is provided by the generalised geometry formulation of the o -shell parameter space presented in appendix C; we also nd a matching between the complex coordinates and the natural parametrisation from generalised geometry. 4 Higher-order deformations The main aim of this paper is to derive the conditions on the moduli when we move from in nitesimal to nite deformations of solutions to the Hull-Strominger system. In other words, we consider higher-order deformations of the elds that parametrise the supersym- metric Minkowski solution. As we have mentioned we only need to consider the F -term relations to understand the moduli space. We show in appendix D that under some reason- able assumptions the D-term conditions do not constrain the moduli problem and should be thought of as gauge xing conditions | we expect this to hold in general. 4.1 The superpotential Let us consider the e ect on the superpotential of a nite deformation of the background elds away from a point on the supersymmetric locus. In other words, we start with a supersymmetric vacuum solution described by a superpotential W which is a functional of the SU(3) structure, H and the bundle. Let us denote the superpotential evaluated at this point by W j0. The vacuum is supersymmetric if both the superpotential and its rst derivative vanish when evaluated on the solution. Now move a nite distance from this solution in parameter space by deforming the background. The superpotential evaluated at this new point is W j = W j0 + W . We have a supersymmetric solution if both W j and its rst derivative vanish at that point in parameter space, which is equivalent to the vanishing of W and W . Let us see how this works out. For clarity of presentation, let us ignore the bundle moduli | we will reinstate these in section 4.3. A nite holomorphic deformation of the parameters gives W = Z X  H + i d! + d(B + i !) ^{ + 1 2 {{ + 1 3! {{{  + d(B + i !) ^  : (4.1) As we are deforming about a supersymmetric point we have H = i(@ @)! and d = 0, so the rst term simpli es and the last term vanishes, giving W = Z X  i @! ^ {{ + d(B + i !) ^  { + 1 2 {{  = Z X h i @! ^ {{ + 2 @x ^ { + @x ^ {{ + @~b ^ { i ; (4.2) { 12 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 where the {{{ term vanishes due to the type of d(B+i !), and we have written the second line in terms of ~b and x, the (0; 2) and (1; 1) parts of the complexi ed hermitian moduli (3.21). As { satis es a graded Leibniz identity, we can rewrite the above as W = 2 Z X  {@x+ 1 2 i {{@! + 1 2 {{@x 1 2 {@~b  ^ = 2 Z X  d ^ @xd + id ^ e ^ @d!ecec + d ^ e ^ @dxe 1 2 d ^ @d~b  ^ : (4.3) Our rst condition for the deformed background to be supersymmetric is W = 0 when evaluated on the solution. In other words, the terms in the brackets in (4.3) should be zero up to a @-exact term. We also need to impose the vanishing of the rst derivative of W . As W is a functional, this amounts to treating it as an action and nding the resulting equations of motion. Varying W , one nds @xa id ^ (@!)dabeb + d ^ @axd d ^ @dxa 1 2 @a~b = 0; (4.4) @d 1 2 [; ]d = 0; (4.5) @{ = 0: (4.6) A few comments are in order. The condition in (4.5) is nothing but the Maurer-Cartan equation for nite deformations of a complex structure. This is somewhat expected as we know solutions to the Hull-Strominger system are manifolds with a complex structure. Notice that we also have a second condition on  in (4.6) which is not usually seen in discussions on the moduli space of complex structures. This condition comes from requiring that the deformed three-form +  is closed and thus holomorphic with respect to the new complex structure | this is stronger than requiring a complex structure alone. Note that this same condition that appears in [40] for Kodaira-Spencer gravity | there it is imposed as a constraint from the outset but one should actually think of it as requiring that the deformed three-form remains d-closed. One could make a change of variables which solves this constraint explicitly by taking { = a+ @b where a is @-harmonic. This may be useful when investigating the quantum theory de ned by the superpotential but we do not use it in what follows. 4.2 A Maurer-Cartan equation from the holomorphic structure The idea now is that these equations can be interpreted as a Maurer-Cartan equation for the deformations. We know the in nitesimal moduli of the Hull-Strominger system are captured by the D-cohomology of the holomorphic structure, so we expect D will be the di erential that appears in such a Maurer-Cartan equation. The other ingredient is a bracket. We introduce the bundle Q0 as the sum of the holomorphic cotangent and tangent bundles: Q0 ' T (1;0)X  T (1;0)X: (4.7) { 13 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 This is the bundle Q de ned in (2.16) with the gauge sector suppressed for now. We write (0; 1)-forms valued in this bundle as y 2 (0;1)(Q0); y = (xaea; ae^a)  x+ : (4.8) The holomorphic structure for the Hull-Strominger system without the bundles is given by an operator D that acts on sections of Q0. We can also introduce a bracket [; ] on forms valued in Q0 and a pairing h; i that traces over the Q0 indices: [; ] : (0;p)(Q0) (0;q)(Q0)! (0;p+q)(Q0); (4.9) h; i : (0;p)(Q0) (0;q)(Q0)! (0;p+q)(X): (4.10) We give explicit expressions for D, the bracket [; ] and the pairing h; i in equa- tions (A.11), (A.13) and (A.16). Using these we can rewrite W , given in (4.3), as W = Z X hy;Dy 1 3 [y; y] @bi ^ ; : (4.11) Note that we have rede ned the (0; 2)-form eld as b = ~b a ^ xa. The equations of motion that follow from varying W can then be written compactly as Dy 1 2 [y; y] 1 2 @b = 0; (4.12) @{ = 0: (4.13) Let us make a few comments. Looking at W in equation (4.11), we see it resembles a Chern-Simons action. More speci cally, the form of the action is that of a holomor- phic Chern-Simons theory for y with a Lagrange multiplier b that enforces a constraint for y. This constraint is the same as the gauge choice that is imposed in Kodaira-Spencer theory [40, 41]. Note that the conventional Chern-Simons action has appeared as a super- potential in other work [42]; we expect a similar analysis could be applied here. Notice also that in nitesimal deformations are captured by Dy = 1 2 @b: (4.14) It follows from this and dH / @@! = 0 that @b is @-closed.6 If the underlying manifold X satis es the @@-lemma or H(0;2)(X) vanishes, @b is @-exact and can be absorbed in a rede nition of the complexi ed Kahler moduli x. We then see that in nitesimally the complexi ed Kahler moduli are counted by H(0;1)(T (1;0)X). 4.3 Including the bundle We now want to include the bundle degrees of freedom in the superpotential | we do this by adding a Chern-Simons term !CS(A) for the gauge connection A: W = Z X dB + !CS(A) + i d!  ^ : (4.15) 6Recall we have turned o the gauge sector in this subsection. { 14 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 The B eld transforms in the usual Green-Schwarz manner so that H is gauge invariant: B = tr(d ^A): (4.16) Consider a shift of the gauge connection by A! A+ : The corresponding change of the Chern-Simons form is !CS = 2 tr(F ^ ) + tr( ^ dA ) + 2 3 tr( ^ ^ ) + d tr( ^A): (4.17) The exact term in this variation combines with the variation of the B eld to give B = B + tr( ^A): (4.18) As we show in appendix B.4, B is a globally de ned two-form so it can be a modulus [14]. Expanding about a supersymmetric point, the extra terms in the variation of the superpotential from the gauge sector are W = Z X !CS ^ ( + { + {{ ) = Z X h (tr( ^ @A ) + 2 3 tr( ^ ^ )) ^ + (2 tr(F ^ ) + tr( ^ @A )) ^ { + d tr( ^A) ^ ({ + 1 2 {{ ) i ; (4.19) where the nal terms combine with b in W to give B = b+ tr( ^ A). Upon replacing b by B in (4.3), the extra terms in the variation of the superpotential are W = Z X  tr( ^ @A ) + 2 3 tr( ^ ^ ) 2d ^Fd ^ ^ d ^ (@A )d  ^ : (4.20) The full expression for W is given by the sum of expressions (4.20) and (4.3) (with b! B). The equations of motion that follow from varying the full superpotential are @xa id ^ (@!)dabeb trFa ^ + d ^ @axd d ^ @dxa + 1 2 tr ^ (@A )a 1 2 @ab = 0; @A + ^ + Fdadza ^ d d ^ (@A )d = 0; @d 1 2 [; ]d = 0; @{ = 0; (4.21) where we have rede ned b to be b = (B + i !)(0;2) a ^ xa: (4.22) Remarkably, the superpotential can still be written in a Chern-Simons fashion as W = Z X hy;Dy 1 3 [y; y] @bi ^ ; (4.23) { 15 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 where y is now (0; 1)-form valued in Q ' T (1;0)XEndV T (1;0)X, and D, [; ] and h; i are now given by expressions (A.19), (A.20) and (A.21). The corresponding equations of motion are Dy 1 2 [y; y] 1 2 @b = 0; (4.24) @{ = 0: (4.25) Together with the vanishing of W , these are the conditions for a supersymmetric Minkowski solution. 4.4 Vanishing of the superpotential We now want to understand the condition W = 0 in more detail. In what follows, we will consider the moduli problem with the gauge sector turned o . Everything we say goes through when we replace the D operator, bracket and pairing with those that include the gauge sector. The deformed vacuum solution is supersymmetric if the equations of motion are satis- ed and the superpotential itself vanishes. As X is assumed to be compact, the superpo- tential vanishes if the terms wedged with are @-exact, that is hy;Dy 1 3 [y; y] @bi = @ ; (4.26) where is an arbitrary (0; 2)-form. Upon substituting the rst equation of motion (4.24) into this expression, it simpli es to 1 3! hy; [y; y]i 1 2 ya@ab = @ : (4.27) Now we use D 2 = 0 to constrain . Taking D of the rst equation of motion (4.24) gives 0 = D 2 y [Dy; y] 1 2 D@b / ea([y; [y; y]]a [y; @b]a + @@ab) = ea  1 3! @ahy; [y; y]i 1 2 @a(y d@db) + @@ab  /  @@b 1 2 @(yd@db) + 1 3! @hy; [y; y]i  ; (4.28) where we have used [y; [y; y]]a = 13!@ahy; [y; y]i and [y; @ab] = 12@a(yd@db).7 We can inte- grate this expression to give k = @b 1 2 yd@db+ 1 3! hy; [y; y]i; (4.29) where k is constant as it is anti-holomorphic and X is compact. Combining this with the vanishing of the superpotential (4.27) gives k = @b+ @ : (4.30) 7These identities are easy to check using the explicit expressions for the bracket and pairing. { 16 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 As is not @ exact, we must have k = 0. We then identify = b up to a @-closed (0; 2)-form. Putting this all together the full set of equations is Dy 1 2 [y; y] 1 2 @b = 0; (4.31) @b 1 2 hy; @bi+ 1 3! hy; [y; y]i = 0; (4.32) @{ = 0: (4.33) These equations are equivalent to the vanishing of the superpotential and its derivative, and so their solutions are a supersymmetric Minkowski vacuum. In other words, solutions (y; b) to these equations are precisely the moduli of the Hull-Strominger system. We pause to make a few comments. First note that these equations contain nitely many powers of y and b; the equations do not give an in nite set of relations. This is somewhat striking | generic deformations of geometric structures do not usually truncate at a given order. In our case the fact that the equations depend on terms up to O(y3) is an indication that there is more structure to the Hull-Strominger system than at rst sight. This extra structure is the existence of an underlying holomorphic Courant algebroid describing the subsector of deformations given by simultaneous deformations of the complex structure and the three-form ux. In deriving these equations and nding third-order equations for the moduli, we might be encouraged to think there is some sort of algebroid underlying the full heterotic system. Indeed, the ten-dimensional heterotic theory has a description in terms of generalised geometry [31, 43, 44] and the Hull-Strominger system can be recast as in terms of holomorphic Courant algebroid [30]. One might wonder if the form of these equations can survive 0 corrections. We derived the equations for the moduli by starting from the superpotential for the four-dimensional theory that one would get by compactifying on a solution to the Hull-Strominger system. Part of the data of such solutions is a complex manifold. A complex manifold admits an SU(3) structure whose torsion is constrained [45]. A special case of such manifolds are those with vanishing torsion so they have SU(3) holonomy and are Calabi-Yau. If the solution to the Hull-Strominger system admits an 0 ! 0 limit, the 0 = 0 solution is simply Calabi- Yau. In this case it is known that the superpotential receives no 0 corrections (to nite order) and so, although the 0-corrected geometry is not longer Calabi-Yau, the tree-level superpotential is exact [22, 46]. This means equations (4.31){(4.33) will be correct even after 0 corrections. It is not known what happens if there is no large-volume Calabi- Yau limit. Up to this point there has been an asymmetry in the way we have treated the vanishing of W and its derivative. In the next section we will see that we can combine these conditions into a single Maurer-Cartan equation for an L3 algebra. 5 Moduli and an L3 algebra So far we have derived the equations that determine the moduli for solutions to the Hull- Strominger system for nite deformations. We will show in this section that these equations { 17 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 can be reinterpreted as the Maurer-Cartan equation for an L3 algebra. At rst sight there is no obvious reason why the deformations of a system as complicated as the heterotic string should be described by such a \nice" algebra structure. However it is not as surprising if one remembers that the data of the Hull-Strominger system is equivalent to a holomorphic Courant algebroid with a holomorphic vector bundle [30]. The L1 structures that govern deformations of Courant algebroids (or Dirac structures) have been found; in particular it is known that the deformation complex of a Dirac structure is isomorphic to a cubic L1 or L3 structure [47{50]. As we review in appendix A.3, an L1 structure is speci ed by a choice of graded vector spaces Yn and multilinear products `k [51, 52]. The idea is that the conditions from the superpotential are most naturally written in terms of an L1 structure that combines the action of D on the moduli elds y and @ on the (0; 2) moduli b. We take the vector spaces Yn to be Yn = (0;n)(Q) (0;n+1)(X); (5.1) so that an element of Y1 is Y = (y; b) where y is a (0; 1)-form valued in Q and b is a (0; 2)-form. Using this notation, we write the multilinear products `k as `1(Y ) :=  Dy 1 2 @b; @b  ; `2(Y; Y ) := ([y; y]; hy; @bi); `3(Y; Y; Y ) := (0;hy; [y; y]i); `k4 := 0: (5.2) We give expressions for the `k where the entries are arbitrary elements of Yn in (A.29). One can check that these products have the correct symmetry properties and obey the L1 relations, which we write in (A.26). Note that this is highly non-trivial and is an indication of the underlying holomorphic Courant algebroid. 5.1 Quasi-isomorphism to a natural holomorphic L3 algebra We brie y remark that these structures have a nice mathematical interpretation: our L3 algebra (Y; `1; `2; `3) is L1 equivalent to the underlying holomorphic algebra. Neglecting the gauge bundle for a moment, we have the sheaf E of D-holomorphic sections of Q0 ' T (1;0)X  T (1;0)X and the sheaf of holomorphic functions OX . These form an L3 algebra, with underlying two-term complex OX @! E ; (5.3) in precisely the same way that a Courant algebroid E together with the real C1 functions form an L3 algebra [53] with two-term complex C1(R) d! E ' T X  TX: (5.4) { 18 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 One can then consider the Dolbeault resolutions of the sheaves E and OX , and extend @ to a morphism between them: 0 OX C1(C) (0;1) (0;2) 0 E Q0 (0;1)(Q0) (0;2)(Q0)  @ @ @  D D D @ @ @ @ (5.5) Our complex Y is then the total complex of the deleted resolution and the di erential `1 is the natural di erential on this (see e.g. [54]). Our construction gives higher `n brackets on Y, providing an L3 algebra structure on the total complex. As (5.5) is simply a resolution of (5.3), this construction essential provides us with a local reformulation of the holomorphic L3 algebra (5.3) in terms of C1 objects. Explicitly, one has a map of complexes (of sheaves) as follows: 0 OX E 0 0 0 C1(C) (E) (0;1) (0;1)(E) (0;2) (0;2)(E) (0;3) @ `1 `1 `1 `1   (5.6) As the cohomology of the total complex is the same as the cohomology of the complex it is resolving, this is a quasi-isomorphism. (One can check this explicitly in our case.) However, the morphism in (5.6) also respects the bracket structure of the L3 algebras on each complex, thus it is a quasi-isomorphism of L3 algebras. We conclude that, in the L1 sense, our L3 algebra (Y; `1; `2; `3) is locally equivalent to the holomorphic algebra (5.3). Including the gauge bundle in this construction is straightforward; one simply replaces the bundle Q0 above with the full holomorphic bundle (C.34) (this also recently appeared in [30]). One nds an essentially identical two-term complex to (5.3) (see [55] for the analogue of (5.4) including the gauge bundle), giving an L3 algebra on the local holomorphic sections. Via the Dolbeault resolution, one sees that this is quasi-isomorphic to our L3 algebra (Y; `1; `2; `3) (now including the gauge bundles) exactly as above. 5.2 An L1 eld equation As explained in [52], there is a natural eld equation that one can write down for a given L1 structure. The constraint on the form of the eld equation is that it is covariant under gauge transformations of the elds Y. In terms of the L1 products, the eld equation is F(Y ) = `1(Y ) 1 2 `2(Y ) 1 3! `3(Y ) + : : : (5.7) For us this expression truncates at third order as `k4 = 0. Remarkably, the L3 eld equation coming from (5.2) reproduces the conditions from the vanishing of the superpotential and its derivative: F(Y ) =  Dy 1 2 @b 1 2 [y; y]; @b 1 2 hy; bi+ 1 3! hy; [y; y]i  : (5.8) { 19 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 In other words, the conditions for a supersymmetric Minkowski vacuum are equivalent to F(Y ) = 0; @{ = 0: (5.9) A particularly nice property of this rewriting is that the L1 structure gives us the gauge transformations of the moduli for free and guarantees that the gauge algebra closes. The gauge transformation of Y by a gauge parameter  = (; ) 2 Y0 is Y = `1() + `2(; Y ) 1 2 `3(; Y; Y ); (5.10) where the higher-order brackets vanish. In general the gauge transformations take eld equations to combinations of eld equations | the eld equations are covariant. If one could construct an action that has the L3 eld equation as its equations of motion, one would expect that action to be invariant under the L3 gauge transformations. In contrast to [52], we have not been able to nd such a candidate action nor do we expect one to exist; this is due to the fact that the supersymmetry conditions are the vanishing of the superpotential and its rst derivative. Note that the superpotential alone is not expected to be invariant under both  and  transformations | we will see in the next section that the superpotential is actually invariant under the  transformations alone which correspond to shifts by @a-exact forms. 6 A reduced L3 algebra and an e ective action In this section we discuss some consequences of the L3 algebra. In particular we comment on how the L3 algebra can be reduced by quotienting by @-exact forms and show that this is equivalent to integrating out b. We also discuss the relation of the moduli (y; b) to the e ective theory one would nd by compactifying on a solution to the Hull-Strominger system. As we have mentioned, the form of the superpotential (4.23) closely resembles that of holomorphic Chern-Simons theory, and is in fact a generalisation of this theory. Holomorphic Chern-Simons theory has several interesting relations with mathematical dis- ciplines such as open and closed topological string theory, knot theory, Donaldson-Thomas invariants and so on, and it would be interesting to look for heterotic generalisations of these relations. This will be the subject of future work. For now, we will restrict ourselves to making some observations about the (semi-) classical e ective action (4.23), and its relation to the lower-dimensional e ective physics. 6.1 Integrating out b We want to integrate out the (0; 2)-form eld b. Looking back at the form of the L3 gauge transformations (5.10), taking  = (0; ) gives ya = 1 2 @a; b = @ 1 2 hy; @i; (6.1) for some (0; 1)-form . One can check that the superpotential (4.23) is invariant under this gauge transformation provided X is compact. From this we see that y is de ned up to { 20 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 @a-exact forms. Notice also that W splits into two pieces: W = Z X hy;Dy 1 3 [y; y]i ^ Z X hy; @bi ^ ; (6.2) where the second term can be written asZ X hy; @bi ^ / Z X b ^ @{ : (6.3) From this we see that b plays the role of a Lagrange multiplier. We shall see below that given certain assumptions about the Hodge diamond of X, speci cally h(2;0) = 0, then the eld b has no associated massless modes. We can then integrate out b, resulting in W [y] = Z X hy;Dy 1 3 [y; y]i ^ ; (6.4) where y now satis es the constraint @{ = 0. We want to think of this functional as an e ective action. Note that for @{ = 0 there is a gauge symmetry of this action ya = @a; (6.5) where  2 0;1(X). We are thus led to de ne ~Q as a reduced sheaf of Q whose sections satisfy the constraint and are de ned up to @a-exact terms: ( ~Q) = f(Q) j @{ = 0; ya  ya + @ag: (6.6) One can check that the brackets on Q are well de ned on ~Q, and that the L3 algebra descends to a di erential graded Lie algebra (DGLA). A very similar sheaf is also considered in the discussion of systems in [56]. Note that this DGLA is L1 quasi-isomorphic to the algebra on (5.1). The second gauge symmetry of (6.4) is a generalisation of the Chern-Simons symmetry. The superpotential is invariant under y = D [y; ]; (6.7) where  2 0( ~Q) satis es @{ = 0. Note that the gauge algebra generated by (5.10) is reducible; a gauge transformation by  = (; ) is trivial if  = @w;  = @w + 1 2 hy; @wi; (6.8) for w 2 0(X). From the e ective superpotential (6.4) we derive the equation of motion Dy 1 2 [y; y] = 0: (6.9) This should be interpreted as an equation on the sheaf ~Q. Note that under ya 7! ya + @a, this equation becomes Dy 1 2 [y; y] = @(@ + yd@d)  0; (6.10) { 21 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 so it is well de ned as an equation on the sheaf ~Q. Recall that we already know the superpotential is invariant under ya = @a. One can show that equation (6.9), together with the condition that the e ective action vanishes, is equivalent to the Maurer-Cartan equations (4.31){(4.33). Indeed, note that (6.9) is equivalent to Dy 1 2 [y; y] = 1 2 @b; (6.11) for some b 2 (0;2)(X). For solutions to this equation, the condition that the action vanishes can then be written as hy; [y; y]i = @-exact, which can be rewritten as 1 3! hy; [y; y]i 1 2 ya@ab = @ ; (6.12) for some 2 (0;2)(X). Here we used that the second term on the left-hand side is @-exact | it integrates to zero against due to the constraint @{ = 0. This then gives the same starting point for our derivation of equations (4.31){(4.33). It is beyond the scope of the present paper to investigate general solutions to (6.9) and W [y] = 0, i.e. integrable deformations of heterotic geometries. We will however make some comments on the couplings derived from (6.4) in the four-dimensional e ective eld theory. From this we make a conjecture about the obstructions that can appear in the Maurer-Cartan equations. 6.2 E ective eld theory and Yukawa couplings Our starting point to derive the e ective physics is the superpotential (4.23), where we have re-introduced the eld b. When dimensionally reducing the theory, it is common practice to split our elds (y; b) into \massless" and \heavy" modes y = y0 + yh; (6.13) b = b0 + bh: (6.14) We imagine performing a formal dimensional reduction of the theory to a four-dimensional Minkowski background where we keep all the massive Kaluza-Klein modes for the time being. The corresponding mass matrix of the reduced theory reads8 V = e K@ @ W@ @W K ; (6.15) where f ; ; : : :g denote holomorphic directions in the parameter space and K is the Kahler potential. Full knowledge of the Kahler potential is not necessary at this point, but the curious reader is referred to [20, 21] for more details. From the form of the mass matrix, it is easy to see that a eld direction is massless if and only if @ @ W = 0 8 ) (@ @ W )j(y;b)=0 = 0 8 ; (6.16) where the eld directions can in principle be massive. 8In principle, there is also a potential coming from D-terms. However, as we show in appendix D, the D-terms can be set to zero by a complexi ed gauge transformation and so they do not lift any moduli. { 22 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 In the end, we are interested in a reduced eld theory of massless modes, where the massive modes have been \integrated out". It is easy to see that the eld direction corre- sponding to b is massive (although need not be an eigenmode of the mass matrix). Indeed, from (6.16) it follows that b0 must satisfy @ab0 = 0; (6.17) and so b is an anti-holomorphic section of (0;2)(X). We restrict ourselves to geometries where this bundle has no sections, in other words H (2;0) @ (X) = 0: (6.18) This is true in particular for Calabi-Yau geometries. It follows that we can integrate out the b eld as far as the e ective theory is concerned, leaving us with the e ective superpo- tential (6.4), where now the Beltrami di erential component  of y satis es @{ = 0 as above. From condition (6.16) it follows that the remaining massless elds y0 then satisfy Dy0 = 0; (6.19) where this should be viewed as an equation in the sheaf ~Q. It is also natural to decompose the symmetry transformations (6.7) in terms of the massless and massive modes. A suggestive decomposition, given the condition (6.19), is the following y0 = D; (6.20) yh = [y0; ] [yh; ]: (6.21) With this decomposition, we see that the massless modes are parametrised by cohomol- ogy classes [y0] 2 H(0;1)D ( ~Q): (6.22) This cohomology is isomorphic to H (0;1) D (Q) for manifolds satisfying either the @@-lemma or H(0;1)(X) = 0 [14, 15]. We give a brief review of this cohomology and its decomposition in into more familiar cohomologies by means of long exact sequences in appendix E. Decomposed in terms of massless and heavy modes, the e ective action now reads W = Z X  hyh; Dy0yhi 1 3 hy0; [y0; y0]i hyh; [y0; y0]i 1 3 hyh; [yh; yh]i  ^ ; (6.23) where we denote Dy0yh = Dyh [y0; yh]: (6.24) We see that the heavy yh modes are the only ones that propagate internally. Note that even though we take the expectation value of yh to vanish, by including internal quantum corrections, we see that the coupling between y0 and yh can generate higher-order couplings of the massless elds. These new couplings are however of quartic order and higher in y0, and are hence non-renormalisable in the e ective eld theory. The only renormalisable { 23 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 coupling we need to worry about from an e ective eld theory point of view is therefore the Yukawa coupling WYuk(y0) = 1 3 Z X hy0; [y0; y0]i ^ : (6.25) This argument is similar to and generalises Witten's standard argument for the gauge sector [22]. Note that in addition to the standard Yukawa couplings between bundle moduli, the Yukawa couplings (6.25) also contain couplings of gravity-gravity type (couplings of de- formations of the geometry) and gravity-bundle type, often referred to as -terms in the literature. It would be interesting to investigate the phenomenological implications of such couplings, but it is beyond the scope of the present paper to do so. Let A 2 H(0;1)( ~Q) denote a set of inequivalent cohomology classes spanning H(0;1)( ~Q), and expand y0 = X A CA A; (6.26) where the CA now correspond to the four-dimensional elds, including in principle moduli and matter elds. The Yukawa couplings then read WYuk = X A;B;C CACBCC Z X h A; [ B; C ]i ^ : (6.27) A massless eld direction A is then truly free if and only if YABC = Z X h A; [ B; C ]i ^ = 0 8 B; C : (6.28) In particular, this is true if [ A; B] = D AB 8 B: (6.29) Note that, starting from the Maurer-Cartan equation (6.9), this is simply the condition for an in nitesimal deformation in the eld direction A to be unobstructed. The e ective eld theory then prompts us to make the following conjecture: when H (2;0) @ (X) = 0,9 the only non-trivial obstructions coming from the Maurer-Cartan equations are given by the constraints (6.29) on the in nitesimal moduli. 7 Conclusions In this paper we have considered nite deformations of the Hull-Strominger system. Start- ing with the four-dimensional N = 1 superpotential, we showed that integrable deforma- tions corresponding to holomorphic directions on the moduli space can be parametrised by solutions of a Maurer-Cartan equation for an L3 algebra, which we described in detail. There are many directions one could follow from this work. Firstly, one might wonder which of the in nitesimal deformations parametrised by H (0;1) D ( ~Q) can be integrated to nite deformations, corresponding to solutions of the L3 Maurer-Cartan equation. In particular, 9So we can integrate out b in the e ective theory. { 24 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 are there some special cases where a generalisation of the Tian-Todorov lemma applies? It would be interesting to apply our formalism to some explicit examples, and in this way work out the spectrum of free elds in the low-energy four-dimensional theory. The superpotential led to a generalisation of holomorphic Chern-Simons theory [57{59] that couples hermitian and complex structure moduli. Following [57] and [40], it seems that one should think of this theory by taking spacetime to be spanned by the anti-holomorphic directions with holomorphic -preserving generalised di eomorphisms playing the role of a gauge group.10 It would be interesting to study this further. In particular, by starting from the superpotential as an e ective action and investigating its quantisation one might hope it de nes a consistent quantum theory (cf. [61, 62]) and gives analogues of Donaldson-Thomas or holomorphic Casson invariants for heterotic geometries. Note that the superpotential is complex in general so the path integral will not be convergent. Such complex path integrals have appeared before in the study of complex Chern-Simons theory [42, 63, 64] where they are understood by analytic continuation. We foresee a similar treatment here. As a step towards a complete understanding of the quantum heterotic moduli space, one could construct a world-sheet AKSZ topological model [65, 66] or a topological string model for the e ective theory similar to Witten's open string model for ordinary Chern-Simons theory [67]. As a guide, one might start by comparing the heterotic moduli space with the spectrum of holomorphic systems and the chiral de Rham complex [56, 68{72]. Several other approaches to the (0; 2) world-sheet have appeared over the years (see [46, 72{79] and references therein). It would be interesting to investigate how these methods connect with the approach outlined in the present paper. These are all interesting aspects which we hope to explore in future publications. One might also consider the moduli space of heterotic compacti cations on more exotic geometries, such as G2 or Spin(7) manifolds [80{82]. In the case of G2 compacti cations, the form of the moduli space is remarkably similar: for example, the in nitesimal deformations are again captured by a cohomology. Despite this there are notable di erences such as the analogue of the bundle Q not appearing as an extension. It would be interesting to investigate the nite deformation algebras in these cases, and in the process identify the corresponding L1 structure. This might give a G2 generalisation of Chern-Simons theory. We have been concerned with nding the honest supersymmetric deformations of solu- tions to the Hull-Strominger system. For this we only needed to consider the superpotential in the four-dimensional theory. Of course, the four-dimensional theory also has a Kahler potential which is important for understanding the physical potential of the e ective the- ory. The Kahler potential and the metric on the moduli space have been worked out in recent publications [20, 21]. One might wonder how these objects appear in our formalism. It seems that the cleanest description of these objects would follow from a proper analysis using generalised geometry. As outlined in appendix C, the N = 1 heterotic structure is 10Generalised di eomorphisms are transformations generated by sections of Q via the Dorfman bracket, whose antisymmetrisation is the Courant bracket. (In the real case, they are simply the di eomorphisms together with gauge transformations of the supergravity elds.) The -preserving condition is just (4.6) which ensures the deformed three-form is d-closed. Another Chern-Simons like theory featuring generalised di eomorphisms as a gauge algebra has recently appeared in [60]. { 25 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 described by an invariant object ~ so the Kahler potential should be given by a functional of this object, similar to the Hitchin functional for SU(3) and G2 structures [83]. Indeed, it seems that a similar story applies to heterotic compacti cations in other dimensions. We hope to make progress on this in a future work. Note that even though we have the invariant object ~ we do not have a natural inte- grability condition for it | the generalised connection is not torsion-free in the heterotic string [43]. Curiously, it appears that when one looks at deformations of this structure there is a nice integrability condition (given by the superpotential). It would be interesting to see if this pattern persists for other generalised geometries. Acknowledgments We would like to thank Bobby Acharya, Dmitri Alekseevsky, Philip Candelas, Jose Figueroa-O'Farrill, Mario Garcia-Fernandez, Marco Gualtieri, Brent Pym, Savdeep Sethi and Dan Waldram for helpful discussions. AA is supported by a Junior Research Fellowship from Merton College, Oxford. XD is supported in part by the EPSRC grant EP/J010790/1. CS-C has been supported by a Seggie Brown Fellowship from the University of Edinburgh. ESS is supported by a grant from the Simons Foundation (#488569, Bobby Acharya). A Conventions In this appendix we set out our conventions and notation. We will use (m;n; : : :) indices to denote real coordinates and (a; b; : : : ; a; b; : : :) to denote complex coordinates on the real six-dimensional manifold X. Using this we can expand, for example, a vector as v = vme^m = v ae^a + v ae^a: (A.1) Our elds are form-valued so, to save space, we often omit the wedge symbol where it will not lead to confusion. Our convention for the contraction of a vector-valued one-form with a p-form is that the vector index is contracted as usual and the form components are wedged. In coordinates, for a vector-valued one-form w and a p-form , we have {w = e m ^ {wm = wn ^ n; (A.2) where m is de ned as m = 1 (p 1)!mn1:::np1e n1:::np1 : (A.3) It follows that {w satis es a Leibniz rule: {w( ^ ) = {w ^  +  ^ {w: (A.4) The interior product of a vector with a one-form is extended to p-vectors and p-forms using the y operation, de ned as uy = 1 p! um1:::mpm1:::mp ; (A.5) where u is a p-vector and  is a p-form. We indicate the p-vector obtained by raising the indices of a p-form with the metric g by a superscript ] | for example (])m1:::mp = gm1n1 : : : gmpnpn1:::np : (A.6) { 26 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 A.1 Heterotic supergravity The Hull-Strominger system follows from setting the supersymmetry variations of the ten- dimensional gravitino , dilatino  and gaugino  to zero. In our conventions these are  M = r+M" = rLCM "+ 1 8 HMNP NP "+O( 02); (A.7)  =  M@M+ 1 12 HMNP MNP  "+O( 02); (A.8)  = 1 2 FMN MN"+O( 02); (A.9) where " is a ten-dimensional Majorana-Weyl spinor and rLC denotes the Levi-Civita connection. A.2 Holomorphic structure Ignoring the gauge sector for the moment, the relevant elds are (0; 1)-forms taking values in Q0 = T (1;0)X  T (1;0)X. We write (0; 1)-forms valued in this bundle as y = (xae a; ae^a) = x+ ; (A.10) where xa and  a are (0; 1)-forms. More generally we will write yp = (xp; p) 2 (0;p)(Q0) | we will often drop the subscript denoting the form degree. The holomorphic structure is de ned by a D operator that is nilpotent. The action of D on y 2 (0;p)(Q0) is (Dy)a = @xa + i(@!)eace c ^ e; (Dy)a = @a: (A.11) One can check that D 2 = 0 follows from dH / @@! = 0. Note also that this convention implies (D@)a = @(@a) = @a@; (A.12) for a form  (not valued in Q0). The bracket [; ] : (0;p)(Q0) (0;q)(Q0)! (0;p+q)(Q0) is [y; y0]a = d ^ @dx0a @dxa ^ 0d 1 2 d ^ @ax0d + 1 2 @axd ^ 0d + 1 2 @a d ^ x0d 1 2 xd ^ @a0d; [y; y0]a = b ^ @b0a @ba ^ 0b: (A.13) The bracket is graded commutative and satis es [yp; yq] = ()1+pq[yq; yp]. Furthermore, D satis es a graded Leibniz identity with the bracket D[yp; y 0 q] = [Dyp; y 0 q] + (1)p[yp; Dy0q]: (A.14) { 27 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 The bracket does not satisfy a graded Jacobi identity. As is the case for a Courant algebroid, the Jacobi identity holds up to a @-exact term. Evaluated for y 2 (0;1)(Q0), one nds [y; [y; y]] = 1 3! @hy; [y; y]i; (A.15) so that only the (1; 0)-form valued component is non-zero. Here we have de ned a pairing between two sections as hy; y0i = d ^ x0d + xd ^ 0d: (A.16) More generally one nds [ym; [y 0 n; y 00 p ]] + (1)m(n+p)[y0n; [y00p ; ym]] + (1)p(m+n)[y00p ; [ym; y0n]] = 1 3! h @ahy; [y0; y00]i+ (1)m(n+p)@ahy0; [y00; y]i+ (1)p(m+n)@ahy00; [y; y0]i i : (A.17) When we include the gauge sector we write sections as y = (xae a; ; ae^a) = x+ +  2 (0;1)(Q); (A.18) where xa and  a are (0; 1)-forms. The D operator, bracket and pairing are extended to include the gauge eld component. The D operator is (Dy)a = @xa + i(@!)eace c ^ e tr(Fa ^ ); (Dy)a = @a; (Dy) = @A + Fb ^ b; (A.19) where the nal component is the gauge eld piece and Fa = Fabdz b. The bracket is [y; y0]a = : : : 1 2 tr ^ (@A 0)a + 1 2 tr(@A )a ^ 0; [y; y0]a = : : : ; [y; y0] = ^ 0 ()1+ 0 0 ^ + b ^ (@A 0)b (@A )b ^ 0b = [ ; 0] + b ^ (@A 0)b (@A )b ^ 0b; (A.20) where we have written only the extra terms that appear in the bracket. Again, the bracket obeys a Jacobi identity up to a @-exact term. The pairing between sections is given by hy; y0i = d ^ x0d + xd ^ 0d + tr( ^ 0): (A.21) A.3 L1 structure We follow [52] for the conventions of an L1 algebra in the \`-picture".11 We start with a graded vector space Y Y = M n Yn; n 2 Z; (A.22) 11We have swapped n! n so that the degree of Yn matches the form degree of y { 28 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 where the Yn are of degree n. The L1 algebra admits multilinear products `1, `2, : : :, where `k has degree 2 k. This means `1 is degree 1, `2 is degree 0, `3 is degree 1, and so on. The products are graded commutative so that for example `2(Y; Y 0) = (1)1+Y Y 0`2(Y 0; Y ); (A.23) where a superscript denotes the degree of Y 2 Yn. More generally we have `k(Y (1); : : : ; Y (k)) = (1)(;Y )`k(Y 1; : : : ; Y k); (A.24) where Y 1 = Y , Y 2 = Y 0, etc. The sign has two contributions: (1) gives a plus if the permutation is even and a minus if the permutation is odd; (;Y ) is determined by Y 1 ^ : : : ^ Y k = (;Y )Y (1) ^ : : : ^ Y (k); (A.25) where Y ^ Y 0 = (1)Y Y 0Y 0 ^ Y . In these conventions, the rst few L1 identities are `1(`1(Y )) = 0; `1(`2(Y; Y 0)) = `2(`1(Y ); Y 0) + (1)Y `2(Y; `1(Y 0)); `1(`3(Y; Y 0; Y 00)) = `3(`1(Y ); Y 0; Y 00) (1)Y `3(Y; `1(Y 0); Y 00) (1)Y+Y 0`3(Y; Y 0; `1(Y 00)) `2(`2(Y; Y 0); Y 00) (1)(Y+Y 0)Y 00`2(`2(Y 00; Y ); Y 0) (1)(Y 0+Y 00)Y `2(`2(Y 0; Y 00); Y ) (A.26) The eld equations and gauge transformations are F(Y ) = 1X n=1 (1)n(n1)=2 (n)! `n(Y n) = `1(Y ) 1 2 `2(Y; Y ) 1 3! `3(Y; Y; Y ) + : : : ; (A.27) Y = `1() + `2(; Y ) 1 2 `3(; Y; Y ) 1 3! `4(; Y; Y; Y ) + : : : ; (A.28) where Y 2 Y1 and  2 Y0. The multilinear products `k for the moduli of the heterotic system are `1(Y ) :=  Dy + 1 2 (1)Y @b; @b  ; `2(Y; Y 0) :=  [y; y0]; 1 2 (hy; @b0i+ (1)1+Y Y 0hy0; @bi)  ; `3(Y; Y 0; Y 00) := 1 3 (1)Y+Y 0+Y 00(0; hy; [y0; y00]i+ (1)Y (Y 0+Y 00)hy0; [y00; y]i + (1)Y 00(Y+Y 0) hy00; [y; y0]i); `k4 := 0: (A.29) { 29 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 B Comments on heterotic ux quantisation In this appendix we comment on ux quantisation in the heterotic string, and its relation to the global nature of the deformation of certain quantities that appear in this paper. Our discussion applies to general ten-dimensional heterotic supergravity and can thus be applied to solutions other than four-dimensional Minkowski compacti cations. Note that we are not saying anything new here; understanding ux quantisation in the heterotic string is still an open problem for the simple reason that H is not d-closed in general. We proceed in steps, beginning with a toy example of the quantisation of the ux of an abelian gauge bundle. We then present a similar treatment of the two-form gerbe as a warm-up for the case of the heterotic gauge elds. We follow [84] for much of the early part of the discussion. B.1 A toy example: the abelian bundle Consider a vector bundle over a manifold X and let A denote an abelian connection with curvature F = dA. Let fU ig denote an open cover of X. We denote the overlaps by U ij = U i \ U j and so on for higher intersections. We assume that the cover is \good" so that the U i and their intersections are contractible. We employ the standard notion for the Cech co-boundary operator where appropriate: for some sheaf F , if fi 2 F(U i) then (@f)ij = fi fj 2 F(U ij), and so on. For the curvature to be well de ned, we require that on U ij we have d(Ai Aj) = 0; (B.1) where Ai and Aj denotes the connections on U i and U j . As U ij is contractible, by the Poincare lemma, we must therefore have (@A)ij = Ai Aj = d ij ; (B.2) on U ij for some zero-forms ij . On triple overlaps U ijk we have d(@ )ijk = d( ij + jk + ki) = (@ 2A)ijk = 0; (B.3) which is often referred to as taking the co-cycle of d ij . It follows that cijk = (@ )ijk are constants: cijk = (@ )ijk = ij + jk + ki 2 R(U ijk)  C1(U ijk;R): (B.4) Clearly from (B.4) we have @c = 0, so the cijk de ne a class of the sheaf cohomology [cijk] 2 H2(X;R) = H2(X;R); (B.5) which represents the two-form ux F = dA. From (B.4) this might look like a trivial co-cycle, but this is deceptive since the class is trivial only if the individual ij can be chosen to be constant. In this case, we see from (B.2) that the connection A can be made global so that the ux is trivial. In this language ux quantisation is the statement that the cijk are in fact integers, cijk 2 Z(U ijk), so that they de ne an integral class [cijk] 2 H2(X;Z) = H2(X;Z) { 30 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 Gauge transformations of the connection do not change the class [cijk] 2 H2(X;R). To see this consider a gauge transformation which preserves the curvature F : A0i = Ai + di; i 2 C1(U i;R): (B.6) This transformation induces a deformation of the ij as 0ij = ij + i j + ij = ij + (@)ij + ij ; (B.7) where ij 2 R(U ij) are constants. Thus c0 = @ 0 = @ + @ = c+ @ is shifted by a Cech co-boundary, and so [c0ijk] = [cijk] 2 H2(X;R) is unchanged. Note that in the case of a quantised ux, the integrality of cijk holds only in a preferred set of gauges for ij under shifts by real constants ij . Now consider a general deformation of the above system. We denote the variations by (: : :): for example A0 = A + A. We x the (integral) cohomology class of the quantised ux so that cijk = 0. This means that we must have (@ )ijk = 0 and thus as C1(R) is acyclic we can nd i 2 C1(U i;R) with  ij = (@)ij = i j . We see the ij can be deformed only by a gauge transformation, as in (B.7). Performing a global gauge transformation A00i = A 0 i di, we nd that in the new gauge we have ~Ai = A 00 i Ai = Ai di; (B.8) so that on U ij we have ~Ai ~Aj = d( ij i + j) = 0: (B.9) We have shown there exists a gauge in which the variation of the connection is a global one-form ~Ai = ~Aj . Note that we could have performed the deformation requiring only that cijk = (@)ijk for ij 2 R(U ij) | this xes the real cohomology class [cijk] 2 H2(X;R). We would then have deduced that  ij = (@)ij + ij , leading to the same gauge transformation of A as above. This would correspond to deforming away from the gauge (choice of ij) in which cijk are explicitly integral, while above we restricted ourselves to the gauge with cijk 2 Z(U ijk). The story becomes more intricate for non-abelian bundles. Recall that a non-abelian connection A on overlaps U ij transforms as Aj = gijAig 1 ij + gijdg 1 ij : (B.10) For the purpose of the present paper we will assume without proof that we can take gij = 0, as in the abelian case. B.2 The two-form gerbe example The case of a two-form gerbe is a direct generalisation of the abelian bundle. The gerbe is speci ed by a set of two-forms Bi 2 2(U i) covering the manifold. The eld strength H = dB is globally de ned. As before, this means that on overlaps U ij we have Bi Bj = dij ; (B.11) { 31 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 for some one-forms ij 2 1(U ij). Again we have (@B)ij = dij on U ij so that (@)ijk is d-closed d(@)ijk = (@ 2B)ijk = 0: (B.12) As U ij is contractible (@)ijk is actually exact (@)ijk = dijk; (B.13) for some ijk 2 C1(U ijk;R). Finally we take the co-cycle of dijkl on quadruple over- laps U ijkl: d(@)ijkl = (@ 2)ijkl = 0: (B.14) It follows that the functions cijkl = (@)ijkl are constants. As before, if the ux is quantised the cijkl can be made integral by appropriate choice of ijk. Indeed, the cijkl represent a co-cycle in the sheaf cohomology [cijkl] 2 H3(X;Z) = H3(X;Z); (B.15) where the class in H3(X;Z) is given by the ux H = dB. Consider deformations of the above system. From the quantisation of the constants cijkl and the acyclicity of C1(R) we have cijkl = (@)ijk = 0 ) ijk = (@)ijk; (B.16) for some ij 2 C1(U ij ;R). This leads immediately to12 (@)ijk = dijk = (@d)ijk ) ij = dij + (@)ij ; (B.17) for some i 2 1(U i). Putting this all together we have Bi Bj = dij = di dj ; (B.18) so that Bi di = Bj dj : (B.19) Looking at these formulae, we see that the variations are forced to take precisely the form needed to constitute a gauge transformation of B and its descendants  and . Thus, in an appropriately chosen gauge, we have that B is a global two-form. B.3 Heterotic ux quantisation We now come to the example relevant for the present paper: the B eld of the heterotic string. Recall that the anomaly cancellation condition reads H = dB + !CS(A); (B.20) 12Note that again this relation would be unchanged if we simply xed the real cohomology class [c] 2 H3(X;R) so that cijkl = (@r)ijkl for some rijk 2 R(U ijk). We would then have ijk = rijk + (@)ijk. { 32 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 where !CS(A) denotes the Chern-Simons three-form and we set 0 4 = 1. A gauge transfor- mation of A of the form A0 = gAg1 + gdg1 (B.21) induces the following transformation of the Chern-Simons form [43] !CS(A 0) = !CS(A) + d tr(g1dg ^A) d(g); (B.22) where d(g) = tr(g dg1g dg1g dg1): (B.23) Hence we have !CS(A 0) = !CS(A) + d!2(g;A); (B.24) where !2(g;A) = tr(g1dg ^A) (g): (B.25) Let us now consider the patching on U ij . From equation (B.20) we get d(Bi Bj) + d!2ij(g;A) = 0; (B.26) where !2ij(g;A) = tr g1ji dgji ^Aj  (gji): (B.27) It follows that Bi Bj + !2ij(g;A) = dij ; (B.28) for some one-forms ij . Taking a co-cycle of this on triple intersections U ijk gives the relation (@!2)ijk = d(@)ijk: (B.29) It can further be shown that (@!2)ijk(g;A) = d! 1 ijk(g); (B.30) for some one-forms !1ijk(g) 2 1(U ijk). The expression for !1ijk(g) is not relevant in the present context, but it can be taken to be independent of the gauge connection A, depending on the transition functions gij . To see this, simply vary !2ij(g;A) with respect to A. We nd that !2ij(g;A) = tr( i ^Ai) tr( j ^Aj); (B.31) where we have de ned A = , which transforms appropriately in the adjoint representa- tion when the transition functions gij are kept constant. From this it is clear that (@!2)ijk(g;A) = 0; (B.32) which shows that we may take !1ijk(g) to be independent of A without loss of generality. Putting together (B.29) and (B.30), we then get the relation (@)ijk !1ijk(g) = dijk; (B.33) { 33 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 for some functions ijk 2 C1(U ijk;R). We again co-cycle this relation on quadruple inter- sections U ijkl to get d !0ijkl(g) + (@)ijkl  = 0; (B.34) where we have used the fact that (@!1)ijkl(g) = d! 0 ijkl(g); (B.35) for some functions !0ijkl(g). We can co-cycle this relation again on quintuple overlaps U ijklm to get d(@!0)ijklm(g) = 0; (B.36) and thus the numbers kijklm = (@! 0)ijklm(g) de ne a class [kijklm] 2 H4(X;R) = H4(X;R): (B.37) This represents the Pontryagin class trF ^ F in the cohomology of the sheaf R. Note however from (B.34) we have for our particular bundle !0ijkl(g) + (@)ijkl = cijkl; (B.38) for some constants cijkl. Computing the co-cycle (@! 0)ijklm(g) we thus nd kijklm = (@c)ijklm; (B.39) and so [kijklm] = 0 2 H4(X;R). This is the sheaf cohomology version of the statement that the bundle in question has a trivial rst Pontryagin class as tr F ^ F = dH is exact. B.4 Deforming the system and a well-de ned global two-form We now consider variations of the above heterotic story. The goal is to show that the heterotic quantisation condition leads us naturally to a global two-form B = B + tr( ^A): (B.40) This is an essential part of the moduli of the main text, where = A is the global one-form variation of the gauge connection. The story is very similar to that of the two gerbe, with some subtleties. Note that as for the abelian bundle, we will assume that we can choose to keep the transition functions gij constant under deformations, that is gij = 0; (B.41) even in the case where the bundle is non-abelian. With this assumption, any deformation of the bundle connection = A can be assumed to be a section of 1(EndV ). We begin by noting that, imposing (B.41), a deformation of (B.38) gives (@)ijkl = cijkl: (B.42) Hence the constants cijkl de ne a sheaf cohomology class in H 3(X;R), even though the original cijkl did not. We will require that this class vanishes [cijkl] = 0 2 H3(X;R), so { 34 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 that the deformation does not produce any new third cohomology. One could think of this as xing the topological data associated to the original numbers cijkl, assuming that they have some notion of \integrality" which cannot be continuously deformed, even though they do not explicitly de ne such a class themselves. Indeed, this condition also seems to agree with the general world-sheet arguments on the heterotic B eld and ux quantisation made in [85], where one thinks of the B eld as a torsor. The lack of the notion of zero in the space of B elds is re ected in our setup by the lack of an explicit cohomology class associated to the cijkl. However, the variations of B about a given starting point do de ne a cohomology class naturally. From this requirement we have (@)ijkl = (@r)ijkl ) ijk = rijk + (@)ijk; (B.43) for some rijk 2 R(U ijk) and ij 2 C1(U ij). Next we take the variation of (B.33), again imposing (B.41), and use the acyclicity of 1(U ijk) to nd (@)ijk = dijk = (@d)ijk ) ij = dij + (@)ij ; (B.44) exactly as for the simple two gerbe case. Finally, we take the variation of (B.28) to obtain Bi di + tr( i ^Ai) = Bj dj + tr( j ^Aj): (B.45) We see we can absorb the d terms via a global gauge transformation of B as before, so that on the overlaps U ij we have Bi = Bj : (B.46) Thus B de nes a global two-form. Note in particular that B is gauge invariant with respect to gauge transformations of the bundle [20]. This follows from the Green-Schwarz mechanism wherein the B eld transforms as B ! B !2(g;A): (B.47) C The o -shell N = 1 parameter space and holomorphicity of In this appendix, we present a description of the space of o -shell scalar eld con gurations that appears when we rewrite the ten-dimensional theory with manifest four-dimensional N = 1 supersymmetry. This rewriting is done in the same spirit as the rewriting of eleven- dimensional supergravity as a four-dimensional N = 8 theory in [86] and the rewriting of type II supergravities as four-dimensional N = 2 theories in [37, 38, 87{89]. Here we focus only on the scalars, which are described as an (in nite-dimensional) space of generalised geometric structures of a type we will specify. We then show how to recover the three- form from the generalised geometric structure and outline how one can see that it is holomorphic on the o -shell eld space, as we claim in section 3. { 35 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 C.1 SU(3) SO(6) structures in the NS-NS sector As a warm up, we consider the common NS-NS sector of ten-dimensional string theories, which is well known to admit a description in the language of SO(10; 10)R+ generalised geometry [90] (see also [91, 92]). When rewriting the theory as a four-dimensional theory, we consider the spacetime to admit a product structure, breaking the ten-dimensional Lorentz symmetry to SO(3; 1)SO(6). The elds which are naturally SO(3; 1) scalars can be packaged as the generalised metric of an SO(6; 6)R+ generalised geometry on the six internal spatial dimensions (see for example [38, 87]). As discussed in [28], the conditions for an N = 1 supersymmetric vacuum can be phrased as the existence of an integrable SU(3)  SO(6) structure on the internal six- dimensional generalised tangent space E ' TX  T X. Such a structure encodes the generalised metric and thus the physical elds. If we require only the presence of an o -shell N = 1 supersymmetry, then one merely restricts the elds to admit a globally de ned spinor, corresponding to a (possibly non-integrable) SU(3) SO(6) structure. The con guration space of o -shell scalar elds that we require in our four-dimensional N = 1 description of the ten-dimensional theory is thus the (in nite-dimensional) space of SU(3) SO(6) structures on the generalised tangent space E. This is the N = 1 NS-NS sector analogue of the discussion of the N = 2 vector- and hyper-multiplet structures given for the full type II theories in [37, 88]. An SU(3) SO(6)  SO(6; 6)R+ structure on E is speci ed by a particular type of complex section ~ of W = L ^3EC; (C.1) where L ' ^6T X is the auxiliary R+ bundle transforming in the 1+1 representation of SO(6; 6) R+ as introduced in [90]. It is convenient to write ~ as ~ = ;  2 (^3EC); (C.2) where  2 (L) is the generalised density with  = pge2 in the coordinate frame (see [90]). In order for ~ to have the correct stabiliser,  must lie in a particular orbit under the action of SO(6; 6) on ^3EC at each point of the manifold X. One can alternatively describe such a generalised structure in terms of a generalised metric, which de nes an SO(6)SO(6) structure on E, which then splits as E ' C+C, so that ^3E ! ^3C+  (^2C+ C) (C+ ^2C) ^3C (C.3) To this, one must add a non-vanishing section of the spin bundle for C+ which we denote { 36 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9  2 (S(C+)). Using local bases E^+m for C+ and E^~m for C de ned by13 E^+m = e^ + m + e + m ie^+mB; E^+~m = e^ ~m e~m ie^~mB; (C.5) one can write an explicit formula for the object   = 1 3! (T mnp)E^+m ^ E^+n ^ E^+p = 1 3! mnpE^+m ^ E^+n ^ E^+p ; (C.6) where is the three-form spinor bilinear of  with itself. Writing the object  in this way, it is guaranteed that it will lie in the correct orbit and its stabiliser is SU(3)  SO(6). Given an N = 2 structure parametrised by two complex pure spinors  2 S(E) L1=2 (see [38, 93]), one can also build an N = 1 structure via the expression ~MNP =  + MNP; (C.7) thus providing a third description of the structure. Having de ned the structure, we will now show how to extract the ordinary complex three-form on the manifold from it. Recall that the generalised tangent space is an extension 0 ! T X ! E ! TX ! 0; (C.8) with the classes of such extensions labelled by the cohomology class of the three-form ux [H] 2 H3(M ;R). The map  is referred to as the anchor map of the Courant algebroid E. This anchor map induces further maps on tensor products of E. In particular we obtain an induced map, which we also label ,  : ^3E ! ^3TX: (C.9) Acting on the bundle L ^3E and using L ' detT X we have  : L ^3E ! detT X ^3TX = ^3T X: (C.10) Thus, applying the anchor map to the generalised structure ~ we obtain an ordinary three- form on the manifold. We now calculate this three-form explicitly. We rst note that (E^m) = e^  m; (C.11) which immediately gives us () = 1 3! mnp(e^m ^ e^n ^ e^p) 2 ^3TXC (C.12) 13These are the split frames of [90], but constructed from an arbitrary local frame for the tangent bundle rather than a vielbein for the ordinary metric. However, the ordinary metric g is still used to lower the indices on the one-form frames em, such that the O(6; 6) metric has components  = g 0 0 g ! (C.4) in these frames. { 37 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 However, we are interested in the object ~ =  and in our frame we have  = p ge2 so that (~) = 1 3! p ge2 mnp(e^m ^ e^n ^ e^p) 2 ^6T X ^3TXC (C.13) Now we use the natural isomorphism ^6T X ^3TX ! ^3T X 1 3! Xmnp(e^m ^ e^n ^ e^p) 7! 1 (3!)2 mnpm0n0p0X m0n0p0(em ^ en ^ ep) (C.14) where m1:::m6(= 1) is the Levi-Civita symbol. The standard metric volume form is then "m1:::m6 = p gm1:::m6 and we nd that under the identi cation (C.14) we have (~) = e2 1 (3!)2 "mnpm0n0p0 m0n0p0(em ^ en ^ ep) = e2(? ) = i e2 : (C.15) To conclude our discussion of the SU(3)SO(6) structure in SO(6; 6)R+ generalised geometry, we note that standard group theoretical arguments give us that the homoge- neous space SO(6; 6) R+ SU(3) SO(6) (C.16) is di eomorphic to the orbit of ~ (at a point in M) under the action of SO(6; 6)  R+ on the complex 220C representation. The homogeneous space (C.16) has a complex structure, with respect to which the element ~ of the 220C is holomorphic, that is the embedding of (C.16) into C220 is a holomorphic map. If we imagine that this complex structure naturally extends to the in nite-dimensional space of SU(3)SO(6) structures on E, then we expect that our generalised three-form ~ will be holomorphic on that space. As the anchor map  is linear and xed by the topology of E, we have that (~) = i e2 = i is also holomorphic on the space of such structures. Note that the decomposition of the Lie algebras appearing here is so(6; 6)! su(3) u(1) so(6) h (3;1)+2  (3;1)2 i R  h (3;6)+1  (3;6)1 i R ; (C.17) so that we can locally parametrise the space (C.16) by exponentiating the action of the complex Lie algebra elements which do not annihilate ~. Taking out the overall scale we have ~0 = ece +  ~; c 2 C 2 (3;1)+2 2 (3;6)+1: (C.18) The parameters c; and then become local complex coordinates on the space (C.16). Note that as SU(3) objects, and carry the same degrees of freedom as the parameters (; (! + i B)(1;1); (! + i B)(0;2)) used to parametrise the space of N = 1 eld con- gurations in sections 3 and 3.2 of the main text, while the parameter c is associated to Kahler transformations. { 38 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 C.2 SU(3) SO(6 + n) structures in heterotic supergravity We can perform a similar analysis to the above in SO(6; 6 + n)R+ generalised geometry for heterotic supergravity [43, 44], where n is the dimension of the gauge group G. In that geometry the generalised tangent space E is an extension of the real Atiyah algebroid A 0 ! C1(g) ! A ! TX ! 0; (C.19) by the cotangent bundle 0 ! T X ! E ! A ! 0; (C.20) with the composition of the maps above still giving an anchor map  : E ! TX. The structure group of the bundle E is then the geometric subgroup (GL(6;R)G)n ((T X g) n ^2T X) of SO(6; 6 + n)  R+, though as usual in generalised geometry we think of E as an SO(6; 6 + n)  R+ vector bundle. The generalised metric de nes an SO(6)  SO(6 + n) structure on E, with a local frame (E^+m; E^~m; E^ ) which can be built from the physical elds (g;B; ;A) (see [43] for details). The presence of a single globally de ned spinor on M breaks the SO(6) factor to SU(3), so that an N = 1 structure is an SU(3) SO(6 + n) structure on E. Using the split frame, one can again write an explicit formula for an object ~ 2 L ^3E de ning such an N = 1 structure: ~ = 1 3! (T mnp)E^+m ^ E^+n ^ E^+p = 1 3! p ge2 mnpE^+m ^ E^+n ^ E^+p : (C.21) As we still have (E^+m) = e^ + m, the argument of appendix C.1 still holds to give us that (~) = i e2 , and similar group theoretical reasoning leads us to conclude that both ~ and e2 = are holomorphic on the coset space SO(6; 6 + n) R+ SU(3) SO(6 + n) (C.22) at each point of X. In this case, the corresponding Lie algebra decomposition reads so(6; 6 + n)! su(3) u(1) so(6 + n) h (3;1)+2  (3;1)2 i R  h (3;6 + n)+1  (3;6 + n)1 i R (C.23) so that we can locally parametrise the orbit of ~ via ~0 = ece +  ~ c 2 C 2 (3;1)+2 2 (3;6 + n)+1: (C.24) As SU(3) G objects, we have the same degrees of freedom as in appendix C.1, but now augmented by a , that is the (0; 1)-form part of the deformation of the gauge eld. These then match the parametrisation of the o -shell N = 1 eld space of section 3 as employed in section 4.3. { 39 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 C.3 The o -shell hermitian structure on V The o -shell parameter space of the Hull-Strominger system is the space of SU(3) struc- tures, B elds and gauge elds for G. On-shell, we know that the gauge bundle V must be a polystable holomorphic bundle. We outline which parts of this structure on the gauge bundle survive o -shell. The gauge group G is a compact unitary real Lie group, which has complex represen- tations. The vector bundle VC is a complex vector bundle with structure group GC, the complexi cation of G. On-shell (on a solution of the Hull-Strominger system), this complex vector bundle has two structures on it: a holomorphic structure and a hermitian structure. The hermitian structure on VC is a reduction of the structure group GC to the compact real form G. This is given by a set of local GC frames for VC on patches of X, which, on the overlaps of patches, are related by transition functions in G  GC only. The holomorphic structure on VC is given by a set of GC frames with respect to which the transition functions are holomorphic functions into the complexi ed group GC. Clearly these two sets of frames are di erent for the simple reason that there are no holomorphic maps into the real group G. From the complex vector bundle VC, we can de ne a real vector bundle V = [VC  V C]R; (C.25) which also admits an action of the complex group GC. The hermitian structure is a positive de nite metric on V which pairs VC and V C, such that in the special frames alluded to above its component matrix takes the canonical form h = 1 2 0   0 ! (C.26) where ; = 1; : : : ; dimC VC are indices for VC and ; = 1; : : : ; dimC VC are indices for V C. O -shell, the manifold X admits an almost complex structure as part of the SU(3) structure. As X is not honestly complex, we lose the holomorphic structure on VC | we cannot de ne holomorphic maps without an integrable complex structure. Nevertheless, physics tells us we have a real connection on VR and that the physical gauge group G is compact and unitary. This means we have the hermitian structure, even o -shell.14 This hermitian structure simply says that VC de nes a real vector bundle V with a compact unitary structure group. The physical connection A is a local section of 1(X; EndV )  1(X; g): (C.27) The almost complex structure de nes a split of this into (1; 0)- and (0; 1)-form parts. As the (1; 0)- and (0; 1)-form summands are intrinsically complex, it is natural to see them as living in the complexi ed Lie algebra, i.e. (1;0)(X; gC) and (0;1)(X; gC), as these are 14As X is only almost complex, one might be tempted to call this an almost hermitian structure to emphasise that the holomorphic structure is not present. { 40 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 de ned via tensor products over the eld C. Naively, the connection has four parts with indices as (Aa) (Aa) (Aa) (Aa) (C.28) However, as the connection is real hermitian, any one of them de nes the rest via com- plex conjugation and multiplication by the hermitian metric on VC. Explicitly, the real condition xes  (Aa)  = (Aa) ;  (Aa)  = (Aa) ; (C.29) while the hermitian condition xes  (Aa)  = h (Aa) h ;  (Aa)  = h (Aa) h : (C.30) Together, these allow us to determine all parts of the connection given only (Aa) ; for example we have (Aa) = h (Aa) h : (C.31) On-shell, the holomorphic structure means that, in each patch, one can choose a GC gauge where (Aa) = 0. O -shell we cannot do this in general, but we can use the previous identities to write any formula purely in terms of (Aa) . For example, the terms appearing in the superpotential involve traces, which simplify using identities like (Aa) (Ab) = (Ab) (Aa) : (C.32) These enable us to write all of the needed expressions using only the objects with indices for VC (eliminating the appearance of objects with indices for V C); for example !CS(A) = tr  A ^ dA+ 2 3 A ^A ^A  = 2  A ^ dA + 2 3 A ^A ^A  : (C.33) so that !CS(A) ^ features only the components (Aa) . Equivalently we could have written this expression purely in terms of V C indices or some combination of the two | this freedom re ects the fact that it is V that appears in the heterotic system, so a split into VC and V C is somewhat arbitrary. The important point is that it is the (0,1) part of the gauge eld that appears. This mirrors the generalised geometry argument from the previous section. Using the freedom to write the Chern-Simons form using only VC indices, one can take the bundle appearing in the deformation complex to be the holomorphic bundle T (1;0)X  EndVC  T (1;0)X: (C.34) which also appears in [30]. { 41 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 D Comments on D-terms In addition to satisfying the F -term conditions derived from the superpotential, a solution should also satisfy the D-term conditions: d e2! ^ ! = 0; (D.1) ! ^ ! ^ F = 0: (D.2) The rst condition is referred to as the conformally balanced condition, while the second condition is the Yang-Mills condition. In this appendix we want to show that these con- ditions impose no extra constraints on the heterotic moduli, given some mild assumptions on the geometry and bundle. This is of course common knowledge from the supergravity point of view [33].15 D.1 Massless deformations We begin by considering compacti cations without bundles. The D-term condition of rel- evance is the conformally balanced condition (D.1). Consider rst a massless deformation of this condition so that y0 = (x0; 0) satis es Dy0 = 0. Our plan is to show that the D-term conditions can be solved order-by-order using the gauge symmetries of x0 so that they do not further constrain the moduli. First we note that deformations of the hermitian form ! and the dilaton  are linked via the SU(3) normalisation condition i 8 ^ = 1 6 e4! ^ ! ^ !: (D.3) Remembering !(0;2) = 0, !(1;1) = ix and !(0;2) = {! i {x, a massless holomorphic variation of this condition gives e40 1 3! ! ^ ! ^ ! = 1 3! ! ^ ! ^ ! i 2 ! ^ ! ^ x0 1 2 ! ^ x0 ^ x0 + i 3! x0 ^ x0 ^ x0: (D.4) Now we expand the massless deformation in terms of a small parameter : y0 =  y(1) +  2 y(2) + : : : (D.5) 0 =  (1) +  2 (2) + : : : (D.6) with a corresponding expansion for x and . At rst and second order the SU(3) normali- sation condition xes (1) = 3 4 i!yx(1); (D.7) (2) = 3 4 i!yx(2) + 1 8 x(1)yx(1): (D.8) 15As we mention in the main text, for an N = 1 supersymmetric theory in four dimensions, supersym- metry breaking is controlled completely by the F -terms when there are no FI parameters. Given a solution to the F -term conditions, one can always make a complex gauge transformation to nd a solution to the D-term conditions on the same orbit. In the heterotic case, this is equivalent to assuming the bundles V and TX are stable. { 42 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 This continues to higher order | the deformation of the dilaton at nth order is xed by the non-primitive part of x(n) and the lower-order elds x(i) for i < n. Now consider a massless holomorphic deformation of the conformally balanced condi- tion. We expand in  and separate into complex type. At O(1) the deformation of the conformally balanced condition reduces to @  i e2! ^ x(1) + (1)e2! ^ !  = 0; (D.9) @  i e2! ^ x(1) + (1)e2! ^ !  = @  e2! ^ {(1)!  : (D.10) Note now that the Gauduchon metric de ned by ~! = e! is balanced as ~!^ ~! is d-closed. We do this as on a hermitian manifold the various Laplace operators for a balanced metric agree on functions [94]: ~df = 2 ~@f = 2 ~@f: (D.11) We also use that the Hodge stars16 on a p-form are related by ? = e(3p)~?: (D.12) Using this we can write the previous equations as @ ~y X(1) = 0; (D.13) @ ~yX(1) = i @ ~y {(1) ~!; (D.14) where we have used the relation between the trace of x(1) and (1) given in (D.7), and we have de ned X(1) = e   x(1) 1 2 ~!yx(1) ~!  : (D.15) The Hodge decomposition for Aeppli cohomology implies that (D.14) and (D.13) determine the (@ + @)-exact part of X(1). Indeed an equivalent set of equations is @ ~y@ ~y X1 = 0; (D.16) @@ ~y X1 = 0; (D.17) @@ ~yX1 = i @@ ~y {1 ~!: (D.18) We now want to argue that these conditions are simply gauge conditions and so do not impose extra conditions on the moduli. Recalling the form of Dy0 from (A.11) we see shifts of x0 by @-exact terms drop out explicitly and that shifts by @-exact terms fall out as we are working modulo @-exact forms.17 A gauge choice for x0 then amounts to a choice of 16We are using the convention for the Hodge star where ^ ? = y vol so that ?! = 1 2 ! ^ ! and ? = i . The dual of a primitive (1; 1)-form p satisfying !y p is ? p = ! ^ p. We also have ? 20 = 20^! where 20 is a (2; 0)-form. This choice satis es ?2 = (1)p on a p-form. The adjoint Dolbeault operators are de ned by @y = ?@?, and we denote the corresponding operators for the Gauduchon metric with a tilde. 17One can also do this calculation with y and b so that the shift by @-exact forms is explicit too. { 43 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 element in @ (0;1)(X) + @ (1;0)(X). We make a simpli cation: let us assume the following cohomologies vanish H (1;0) @ (X) = H (2;0) @ (X) = H (0;1) @ (X) = H (0;2) @ (X) = 0; (D.19) so that we get a Hodge decomposition of the space of (@ + @)-exact forms as @ (0;1) + @ (1;0) = @@ (0;0)  @@~y (2;0)  @@~y (0;2): (D.20) Note that it is important that we include the dilaton degrees of freedom: shifts of x(1) are generically not primitive and so they will change the SU(3) normalisation condition, but we can compensate for this by shifting (1) (which does not appear explicitly in the rst-order conformally balanced condition). We start by shifting x(1) by @@(1), where (1) is a function. A short calculation shows that equation (D.16) becomes @ ~y@ ~y  e  x(1) 1 2 ~!yx(1) ~!  = @ ~y@ ~y e@@(1)  + 1 2 ~@  e ~@(1)  : (D.21) One can check that the operator acting on (1) is a positive semi-de nite self-adjoint elliptic operator whose image is given by non-constant functions. This means that (D.21) can always be solved by an appropriate choice of (1). This xes the @@-gauge symmetry of x(1). Next consider a shift x(1) by @@~y (1),18 where (1) is a (2; 0)-form. A short calculation shows that (D.17) becomes @@ ~y  e  x(1) 1 2 ~!yx(1) ~!  = @@ ~y (e@@~y (1)): (D.22) Again, one can check that the operator acting on (1) is positive semi-de nite, self-adjoint and elliptic so that (D.1) can always be solved for by a choice of (1). This xes the @@ ~y-gauge symmetry of x(1). Finally consider a shift of x(1) by @@ ~y (1), where (1) is a (0; 2)-form. A short calcu- lation shows that (D.18) becomes @@ ~y  e  x(1) 1 2 ~!yx(1) ~!  i @@~y({(1) ~!) = @@ ~y  e@@ ~y (1)  : (D.23) As {(1)! is a (0; 2)-form and we assume H (0;2)(X) vanishes, @@ ~y {(1)! is actually @@ ~y-exact. Thus when we shift x(1), this equation can be solved providing the operator acting on the gauge parameter is elliptic and positive semi-de nite as before | it is simple to check that this is the case. This xes the @@ ~y -gauge symmetry of x(1). What happens at higher orders in ? At second order we have @ ~y (X(2) +A(2)) = 0; (D.24) @ ~y(X(2) +A(2)) = i @ ~y {(2) ~! + @ ~y B(2); (D.25) 18One should think of this x(1) as already gauge transformed to solve the previous condition. { 44 { J H E P 1 0 ( 2 0 1 8 ) 1 7 9 where X(2)(x(2)) = e (x(2) 1 2 ~!yx(2) ~!): (D.26) A(2)(x(1)) = i e   3 8 (!yx(1))2! + 1 4 x(1)yx(1) ! + 1 2 !yx(1) x(1) + ?(x(1) ^ x(1))  ; (D.27) B(2)(x(1); (1)) = e   1 2 !yx(1) {(1)! + {(1)x(1) + ?(x(1) ^ {(1)!)  : (D.28) We see that X(2) depends only on the second-order correction to x0 while A(2) and B(2) are xed by the rst-order terms (which should be thought of as gauge transformed to solve the rst-order conditions). Again (D.24) and (D.25) are equivalent to @ ~y@ ~y (X(2) +A(2)) = 0; (D.29) @@ ~y (X(2) +A(2)) = 0; (D.30) @@ ~y(X(2) +A(2)) = i @@ ~y {(2) ~! + @@ ~y B(2): (D.31) As X(2) is a function of x(2) alone, we can perform gauge transformations of x(2) without a ecting A(2) and B(2). Generically these gauge transformations will break the SU(3) normalisation condition, but we can always shift (2) to compensate for this (which is what we have implicitly done by eliminating (2) from the equations). An analogous argument to the one we gave previously then shows that we can always solve these conditions using the gauge freedom of x(2). From this it is simple to see that this process can be continued to all orders. The conformally balanced condition at order n is a set of equations for x(n) with x(i