This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

aainstitutetext: Department of Physics, Federal University of Espírito Santo,
29075-900, Vitória, Brazil
bbinstitutetext: Institute for Nuclear Research and Nuclear Energy,
Bulgarian Academy of Sciences, 1784 Sofia, Bulgaria

Renormalization of Twisted Ramond Fields in D1-D5 SCFT2

A. A. Lima,111Corresponding author. a    G. M. Sotkov b    M. Stanishkov [email protected] [email protected] [email protected]
Abstract

We explore the nn-twisted Ramond sector of the deformed two-dimensional 𝒩=(4,4)\mathcal{N}=(4,4) superconformal (T4)N/SN(T^{4})^{N}/S_{N} orbifold theory, describing bound states of D1-D5 brane system in type IIB superstring. We derive the large-NN limit of the four-point function of two R-charged twisted Ramond fields and two marginal deformation operators at the free orbifold point. Specific short-distance limits of this function provide several structure constants, the OPE fusion rules and the conformal dimensions of a few non-BPS operators. The second order correction (in the deformation parameter) to the two-point function of the Ramond fields, defined as double integrals over this four-point function, turns out to be UV-divergent, requiring an appropriate renormalization of the fields. We calculate the corrections to the conformal dimensions of the twisted Ramond ground states at the large-NN limit. The same integral yields the first-order deviation from zero of the structure constant of the three-point function of two Ramond fields and one deformation operator. Similar results concerning the correction to the two-point function of bare twist operators and their renormalization are also obtained.

Keywords:
Symmetric product orbifold of 𝒩=4\mathcal{N}=4 SCFT, marginal deformations, twisted Ramond fields, correlation functions, anomalous dimensions.

1 Introduction

In the near-horizon decoupling limit Maldacena:1997re of the type-IIB supergravity description of bound states of N1N_{1} D1- and N5N_{5} D5-branes, the asymptotic geometry becomes AdS3×S3×T4AdS_{3}\times S^{3}\times T^{4}, with large Ramond-Ramond charges Maldacena:1998bw ; Seiberg:1999xz , from which one can reconstruct its holographic dual SCFT2; see David:2002wn ; Mathur:2005zp ; Skenderis:2008qn for reviews. There is strong indication that this D1-D5 SCFT2 flows in the infrared to a free field theory whose sigma model is (T4)N/SN(T^{4})^{N}/S_{N}, an orbifold of T4T^{4} by the symmetric group SNS_{N}, with N=N1N5N=N_{1}N_{5}, while the supergravity description is obtained by moving in moduli space with a deformation away from this ‘free orbifold point’. Gravitational solutions, which include the Strominger-Vafa black hole Strominger:1996sh and fuzzball geometries Lunin:2001jy ; Mathur:2005zp ; Kanitscheider:2007wq ; Kanitscheider:2006zf ; Skenderis:2008qn ; Mathur:2018tib , are all dual to states in the Ramond sector of the SCFT2, and one can find correspondences between the geometries and Ramond states in many cases where the latter are BPS-protected from renormalization. Extensive research has achieved considerable progress in the understanding of the free orbifold and its deformation, as well as in the construction of ‘superstratum’ geometries corresponding to the microscopic picture Lunin:2000yv ; Lunin:2001pw ; Balasubramanian:2005qu ; Avery:2009tu ; Pakman:2009ab ; Pakman:2009mi ; Burrington:2012yq ; Bena:2013dka ; Carson:2014ena ; Carson:2015ohj ; Fitzpatrick:2016ive ; Burrington:2017jhh ; Galliani:2017jlg ; Bombini:2017sge ; Tormo:2018fnt ; Eberhardt:2018ouy ; Bena:2019azk ; Dei:2019osr ; Eberhardt:2019ywk ; Giusto:2018ovt ; Martinec:2019wzw ; Hampton:2019csz ; Warner:2019jll ; Dei:2019iym ; Galliani:2016cai ; Bombini:2019vnc ; Giusto:2020mup . Nevertheless, the description of the dynamics of the deformed SCFT2 is still not fully understood. One of the open problems concerns the selection rules separating protected states from “lifted” ones, whose conformal data flow in the deformed theory after renormalization Guo:2019pzk ; Keller:2019suk ; Keller:2019yrr ; Guo:2019ady ; Belin:2019rba ; Guo:2020gxm .

The present paper investigates the effects on the conformal properties of twisted ground states in the 𝒩=(4,4)\mathcal{N}=(4,4) orbifold SCFT2 when the theory is deformed by a marginal scalar modulus operator λO[2](int)\lambda O^{(\mathrm{int})}_{[2]} Avery:2010er ; Avery:2010hs ; Carson:2014ena ; Carson:2015ohj ; Carson:2016uwf . The first-order correction, in powers of λ\lambda, of the two-point function of a ground state is known to vanish. Our main result is an explicit derivation of the finite part of the second-order correction to two-point functions of nn-twisted primary operators 𝒪[n]\mathscr{O}_{[n]}, by eliminating the UV divergences with an appropriate renormalization of the fields. As a consequence, the scaling dimension Δn𝒪(0)\Delta^{\mathscr{O}}_{n}(0) in the free orbifold point flows with λ\lambda according to

Δn𝒪(λ)=Δn𝒪(0)+π2λ2|J𝒪(n)|,\Delta_{n}^{\mathscr{O}}(\lambda)=\Delta^{\mathscr{O}}_{n}(0)+\frac{\pi}{2}\lambda^{2}|J_{\mathscr{O}}(n)|, (1)

where J𝒪(n)J_{\mathscr{O}}(n) is a regularized integral defined in Sect.6 below. We will be mostly interested in two specific operators: the bare twist field σ[n]\sigma_{[n]}, and, more importantly, the twisted R-charged Ramond fields R[n]±R^{\pm}_{[n]}, with bare (holomorphic) conformal weight h=n4h=\frac{n}{4}. In both cases, our main result — the lifting of the conformal dimensions — holds for n<Nn<N. Ramond fields with n=Nn=N and n=2n=2, i.e. with maximal and minimal twist, are protected at leading order in the large-NN approximation. Bare twist fields with n=Nn=N are also protected at leading order. For the particular case of twisted Ramond fields, this has been recently reported in our short letter Lima:2020boh .

The main ingredient in the calculation of second-order corrections to the two-point functions is finding an explicit analytic expression for the four-point function

𝒪[n](z1,z¯1)O[2](int)(z2,z¯2)O[2](int)(z3,z¯3)𝒪[n](z4,z¯4)\big{\langle}\mathscr{O}_{[n]}^{\dagger}(z_{1},\bar{z}_{1})O^{(\mathrm{int})}_{[2]}(z_{2},\bar{z}_{2})O^{(\mathrm{int})}_{[2]}(z_{3},\bar{z}_{3})\mathscr{O}_{[n]}(z_{4},\bar{z}_{4})\big{\rangle} (2)

for the fields 𝒪[n]\mathscr{O}_{[n]} we are interested in. We will present a detailed derivation of the large-NN approximation of the corresponding functions, by applying covering surface techniques Lunin:2000yv ; Lunin:2001pw combined with the ‘stress-tensor method’ Dixon:1985jw . Our computation of the leading term in the 1/N1/N expansion of the connected part of (2), takes into account only the terms contributing to genus zero surfaces, i.e. we use the well known map Arutyunov:1997gi ; Arutyunov:1997gt between the “base” branched sphere to its genus-zero covering surface Lunin:2000yv . The alliance of the covering surface with the stress-tensor method emphasizes some interesting mathematical properties of the correlation functions, and their relation to Hurwitz theory Pakman:2009ab ; Pakman:2009zz ; Pakman:2009mi .

Corrections to the anomalous dimensions at second-order follow from the integral of (2) over the positions of the interaction operators. The analogous integral in the case where there are NS chiral fields at z1z_{1} and z4z_{4} has been computed in Pakman:2009mi , and shown to vanish, as expected for protected operators which should not renormalize. For the twisted Ramond fields and the twist operators, however, the integrand has a more complicated structure, with one more branch cut, and without appropriate regularization the integrals are divergent. In order to define and evaluate their finite parts, we have elaborated a regularization procedure and a specific renormalization scheme for the fields in the deformed SCFT2. Our starting point is the observation that the integrals we are interested in can be put in a form studied by Dotsenko and Fateev Dotsenko:1984nm ; Dotsenko:1984ad ; dotsenko1988lectures in a different context, as integral representation of the conformal blocks of primary fields (curiously; not of their integrals) in the c<1c<1 series222See Refs.Mussardo:1987eq ; Mussardo:1987ab ; Mussardo:1988av for the extension of the Dotsenko-Fateev integral representation to the Ramond and twisted sectors of the 𝒩=1\mathcal{N}=1 and 𝒩=2\mathcal{N}=2 supersymmetric minimal models. of minimal CFT2 models. While, in the one hand, they can be formally written as specific contour integrals in the complex plane — with the contours ensuring a series of algebraic properties — on the other hand these integrals can be represented by four ‘canonical functions’ which are analytic in their parameters, even in cases where the integral itself diverges. Thus, by analytic continuation, the canonical functions give a regularized result for the desired integrals of the four-point functions (2). When applied to the parameters of NS chirals, this procedure gives a vanishing result, as expected; but when applied to the Ramond and twisted fields, we find finite, non-vanishing corrections to the conformal dimensions. The analytic expressions for the renormalized conformal dimensions of R[n]±R^{\pm}_{[n]} and σ[n]\sigma_{[n]} is one of the most important results of this paper. As a byproduct of the computation of the integrals, we can also present the first-order correction to the structure constants R[n]()O[2](int)(1)R[n]+(0)\langle R_{[n]}^{-}(\infty)O^{(\mathrm{int})}_{[2]}(1)R_{[n]}^{+}(0)\rangle, and σ[n]()O[2](int)(1)σ[n](0)\langle\sigma_{[n]}(\infty)O^{(\mathrm{int})}_{[2]}(1)\sigma_{[n]}(0)\rangle, which do not vanish in the deformed SCFT2. It is worthwhile to mention the recent use of similar methods to the renormalization of certain composite Ramond fields, for example R[n]+(z)R[m](z)R_{[n]}^{+}(z)R^{-}_{[m]}(z) Lima:2020nnx . In the composite case, an important consequence of the renormalization procedure is the existence of a condition, namely m+n=Nm+n=N, selecting a class of protected (non-renormalized) states. The remaining states, with n+m<Nn+m<N, are lifted: their renormalized conformal dimensions flow with λ\lambda, and are given by the sum of the second-order corrections (1) for each one of the constituents, i.e. ΔnR(λ)+ΔmR(λ)\Delta_{n}^{R}(\lambda)+\Delta_{m}^{R}(\lambda). What distinguishes the protected Ramond fields from the lifted ones is that the former have conformal weight h=124corb=14Nh=\frac{1}{24}c_{orb}=\frac{1}{4}N, i.e. they are Ramond ground states of the full orbifold theory; in contrast, the fields with h=14nh=\frac{1}{4}n, n<Nn<N are Ramond ground states only of the nn-twisted sector, or of the ‘nn-wound component string’ in the familiar description of the D1-D5 SCFT in terms of effective stings.

Another contribution of the present paper is the analysis of short-distance limits of the four-point function (2). In the limits where operators coincide, u0,1,u\to 0,1,\infty, we are able to derive several structure constants, the OPE fusion rules and the conformal dimensions of some non-BPS operators. These OPE data add to the description of the Ramond and twisted sectors of the free-orbifold point. Our results for the non-BPS fields are consistent with what is known about the chiral NS and twisted sectors Pakman:2009mi ; Burrington:2017jhh . They are also in agreement with the recently conjectured universality of OPEs of certain chiral fields and the deformation operator in the large-NN limit Burrington:2018upk ; deBeer:2019ioe , and represent an extension of these results for all other sectors of the free orbifold theory. In particular, we find that the OPE algebra of the deformation operator and the Ramond fields includes a set of R-charged twisted non-BPS operators Ym±Y^{\pm}_{m}, appearing in the OPEs O[2](int)(z,z¯)Rn±(0)O^{(\mathrm{int})}_{[2]}(z,\bar{z})R^{\pm}_{n}(0). Similarly, the algebra of O[2](int)O^{(\mathrm{int})}_{[2]} and σn\sigma_{n} includes new twisted operators 𝒴m\mathcal{Y}_{m}. We have calculated the dimensions of these operators, as well as the values of structure constants such as Rn±()O2(int)(1)Yn±1±(0)\langle R^{\pm}_{n}(\infty)O^{(\mathrm{int})}_{2}(1)Y^{\pm}_{n\pm 1}(0)\rangle. Applying the fractional spectral flows of Ref.deBeer:2019ioe with ξ=n/(n+1)\xi=n/(n+1), we find that our results for the twisted Ramond fields’ OPEs are in complete correspondence with those obtained from OPEs in the NS sector resulting in specific non-BPS NS fields.

The structure of the paper is as follows. In Sects.2 and 3, we fix our notations by defining first the free orbifold SCFT2, and then its deformation away from the free orbifold point; we also review some key features of conformal perturbation theory used later. In Sect.4, we give a detailed calculation of the four-point functions involving Ramond and bare twist fields, necessary for the second-order correction of the two-point functions. In Sect.5, we investigate certain short-distance limits of the four-point function in order to extract OPE fusion rules, conformal weights and structure constants of several operators in the free-orbifold point. In Sect.6, we return to conformal perturbation theory, with a detailed study of the regularization and the final computation of integrals resulting in the change of the conformal weights of R[n]±R^{\pm}_{[n]} and σ[n]\sigma_{[n]}; we also explain how the renormalization scheme can be extended to a generic primary field 𝒪[n]\mathscr{O}_{[n]}; we also comment on the spectral flow between the fields R[n]±R^{\pm}_{[n]} and NS chiral operators in the free theory, and how it is “broken” after the deformation. In Sect.7 we present a compact summary of our results, together with a short discussion of a few open problems and the eventual consequences of the continuous (λ\lambda-dependent) conformal dimensions of the renormalized twisted Ramond fields for their geometric bulk counterparts. Some auxiliary topics are left for the appendices.

2 D1-D5 SCFT2 and (T4)N/SN(T^{4})^{N}/S_{N} orbifold

The ‘free orbifold point’ of the D1-D5 system is the SCFT2 with central charge (corb,c~orb)=(6N,6N)(c_{orb},\tilde{c}_{orb})=(6N,6N), obtained by taking NN copies of the free 𝒩=(4,4)\mathcal{N}=(4,4) SCFT2, identified under the symmetric group SNS_{N}, with target space (T4)N/SN(T^{4})^{N}/S_{N}. The 𝒩=(4,4)\mathcal{N}=(4,4) superconformal algebra of the ‘seed theory’ has central charge (c,c~)=(6,6)(c,\tilde{c})=(6,6), R-symmetry group SU(2)L×SU(2)R\mathrm{SU(2)}_{L}\times\mathrm{SU(2)}_{R}, and ‘internal’ group SO(4)I=SU(2)1×SU(2)2\mathrm{SO(4)_{I}}=\mathrm{SU(2)}_{1}\times\mathrm{SU(2)}_{2} corresponding to the torus T4T^{4} of target space. We work on the complex plane, with coordinates z,z¯z,\bar{z}.

The unitary representations of the holomorphic 𝒩=4\mathcal{N}=4 algebra are characterized by three numbers {h,j3,𝔧3\{h,j^{3},\mathfrak{j}^{3}}, respectively the conformal weight and the semi-integer charges under the R-current J3(z)J^{3}(z) of SU(2)L, and a current 𝔍3(z)\mathfrak{J}^{3}(z) of the global SU(2)1. Similar numbers {h~,j~3,𝔧~3}\{\tilde{h},\tilde{j}^{3},\tilde{\mathfrak{j}}^{3}\} characterize the anti-holomorphic sector with SU(2)R and SU(2)2 groups.

The theory can be realized in terms of free bosons XA˙A(z,z¯)X^{\dot{A}A}(z,\bar{z}) and free fermions ψαA˙(z)\psi^{\alpha\dot{A}}(z), ψ~α˙A˙(z¯)\tilde{\psi}^{\dot{\alpha}\dot{A}}(\bar{z}), whereas the stress-tensor, the R-current and the super-current are expressed as

T(z)\displaystyle T(z) =14ϵA˙B˙ϵABXA˙AXB˙B+14ϵA˙B˙ϵαβψαA˙ψβB˙\displaystyle=\tfrac{1}{4}\epsilon_{\dot{A}\dot{B}}\epsilon_{AB}\partial X^{\dot{A}A}\partial X^{\dot{B}B}+\tfrac{1}{4}\epsilon_{\dot{A}\dot{B}}\epsilon_{\alpha\beta}\psi^{\alpha\dot{A}}\partial\psi^{\beta\dot{B}} (3a)
Ja(z)\displaystyle J^{a}(z) =14ϵA˙B˙ϵαβψαA˙[σa]βψγB˙γ\displaystyle=\tfrac{1}{4}\epsilon_{\dot{A}\dot{B}}\epsilon_{\alpha\beta}\psi^{\alpha\dot{A}}[\sigma^{*a}]^{\beta}{}_{\gamma}\psi^{\gamma\dot{B}} (3b)
GαA(z)\displaystyle G^{\alpha A}(z) =ϵA˙B˙ψαA˙XB˙A\displaystyle=\epsilon_{\dot{A}\dot{B}}\psi^{\alpha\dot{A}}\partial X^{\dot{B}A} (3c)

with similar expressions for the anti-holomorphic sector. Conventions for SU(2) indices are given in Appendix A. The complex bosons and the complex fermions obey reality conditions (158), and can be written in terms of real bosons and fermions Xi(z,z¯)X_{i}(z,\bar{z}), ψi(z)\psi_{i}(z) and ψ~i(z¯)\tilde{\psi}_{i}(\bar{z}), i=1,2,3,4i=1,2,3,4; see (157). The fermions can be described in terms of chiral scalar bosons ϕr(z)\phi_{r}(z) and ϕ~r(z¯)\tilde{\phi}_{r}(\bar{z}), with r=1,2r=1,2. In the holomorphic sector,

[ψ+1˙(z)ψ1˙(z)]=[eiϕ2(z)eiϕ1(z)],[ψ+2˙(z)ψ2˙(z)]=[eiϕ1(z)eiϕ2(z)].\begin{bmatrix}\psi^{+\dot{1}}(z)\\ \psi^{-\dot{1}}(z)\end{bmatrix}=\begin{bmatrix}e^{-i\phi_{2}(z)}\\ e^{-i\phi_{1}(z)}\end{bmatrix}\ ,\qquad\begin{bmatrix}\psi^{+\dot{2}}(z)\\ \psi^{-\dot{2}}(z)\end{bmatrix}=\begin{bmatrix}e^{i\phi_{1}(z)}\\ -e^{i\phi_{2}(z)}\end{bmatrix}. (4)

Every exponential should be understood to be normal-ordered (and we ignore cocycles). The stress-tensor (3a) can be written in the completely bosonic form,333Normal ordering of two operators 𝒜1\mathscr{A}_{1} and 𝒜2\mathscr{A}_{2} is defined by :𝒜1(z)𝒜2(z):limε0[𝒜1(z+ε)𝒜2(z)𝒜1(z+ε)𝒜2(z)].\,\colon\!\mathscr{A}_{1}(z)\mathscr{A}_{2}(z)\!\colon\,\equiv\lim_{\varepsilon\to 0}\Big{[}\mathscr{A}_{1}(z+\varepsilon)\mathscr{A}_{2}(z)-\langle\mathscr{A}_{1}(z+\varepsilon)\mathscr{A}_{2}(z)\rangle\Big{]}.

T(z)=12limzz[i=1412Xi(z)Xi(z)+r=12ϕr(z)ϕr(z)+c(zz)2].T(z)=-\frac{1}{2}\lim_{z^{\prime}\to z}\Bigg{[}\sum_{i=1}^{4}\frac{1}{2}\partial X_{i}(z)\partial X_{i}(z^{\prime})+\sum_{r=1}^{2}\partial\phi_{r}(z)\partial\phi_{r}(z^{\prime})+\frac{c}{(z-z^{\prime})^{2}}\Bigg{]}. (5)

Bosons are assumed to be periodic, so e.g. XA˙A(e2πiz)=XA˙A(z)X^{\dot{A}A}(e^{2\pi i}z)=X^{\dot{A}A}(z). Fermions can have Neveu-Schwarz or Ramond boundary conditions on \mathbb{C}. The Ramond sector has a collection of degenerate vacua with holomorphic dimension h=c24=14h=\frac{c}{24}=\frac{1}{4}, and different charges under the global and R-symmetry SU(2) groups. The set of Ramond vacua can be obtained from the NS vacuum by the action of spin fields, conveniently realized as exponentials, e.g. for the SU(2) doublet Sα(z)S^{\alpha}(z),

S±(z)\displaystyle S^{\pm}(z) =e±i2[ϕ1(z)ϕ2(z)].\displaystyle=e^{\pm\frac{i}{2}[\phi_{1}(z)-\phi_{2}(z)]}. (6)

To construct the orbifold (T4)N/SN(T^{4})^{N}/S_{N}, one makes NN copies of the free SCFT and identifies them under the action of SNS_{N}; more explicitly, we take the NN-fold tensor product (NT4)/SN(\otimes^{N}T^{4})/S_{N}, and label operators 𝒪I\mathscr{O}_{I} in each copy by an index I=1,,NI=1,\cdots,N. The energy tensor becomes

T(z)=12limzzI=1N[i=1412XiI(z)XiI(z)+r=12ϕrI(z)ϕrI(z)+c(zz)2].T(z)=-\frac{1}{2}\lim_{z^{\prime}\to z}\sum_{I=1}^{N}\Bigg{[}\sum_{i=1}^{4}\frac{1}{2}\partial X_{iI}(z)\partial X_{iI}(z^{\prime})+\sum_{r=1}^{2}\partial\phi_{rI}(z)\partial\phi_{rI}(z^{\prime})+\frac{c}{(z-z^{\prime})^{2}}\Bigg{]}. (7)

and the total central charge is corb=Nc=6Nc_{\mathrm{orb}}=Nc=6N.

Permutations of the copies can be realized by the insertion of twist operators σg(z)\sigma_{g}(z), gSNg\in S_{N}, which give a representation of SNS_{N}, and act on the other operators by twisting their boundary conditions Dixon:1986qv ,

𝒪I(e2πiz)σg(0)=𝒪g(I)(z)σg(0).\mathscr{O}_{I}(e^{2\pi i}z)\sigma_{g}(0)=\mathscr{O}_{g(I)}(z)\sigma_{g}(0). (8)

We are going to consider only cyclic twists, which form the building blocks of the Hilbert space of the orbifold theory Dijkgraaf:1996xw . So, denoting by (n)(n) a generic cycle of length nn, we consider g=(1)Nn(n)(n)g=(1)^{N-n}(n)\cong(n), leaving the trivial cycles implicit. We denote by σn\sigma_{n} the twist operator for a generic cycle of length nn; they cyclically permute the nn copies of the fields appearing in the cycle (n)(n), while leaving the remaining copies invariant. The conjugacy class is represented by the orbit-invariant combination

σ[n]1𝒮n(N)hSNσh1(n)h\sigma_{[n]}\equiv\frac{1}{\mathscr{S}_{n}(N)}\sum_{h\in S_{N}}\sigma_{h^{-1}(n)h} (9)

where the representing cycle can be taken to be (n)=(1n)(n)=(1\cdots n), and the combinatorial factor 𝒮n(N)\mathscr{S}_{n}(N) makes the two-point function normalized, i.e.

σ[n](z1,z¯1)σ[m](z2,z¯2)=δmn|z1z2|2Δn.\big{\langle}\sigma_{[n]}(z_{1},\bar{z}_{1})\sigma_{[m]}(z_{2},\bar{z}_{2})\big{\rangle}=\frac{\delta_{mn}}{|z_{1}-z_{2}|^{2\Delta_{n}}}. (10)

The well-known (total) conformal dimension of a twist σn(z,z¯)\sigma_{n}(z,\bar{z}) is Lunin:2001pw ; Dixon:1986qv

Δnσ=hnσ+h~nσ,hnσ=14(n1n)=h~nσ.\Delta^{\sigma}_{n}=h^{\sigma}_{n}+\tilde{h}^{\sigma}_{n},\qquad h_{n}^{\sigma}=\frac{1}{4}\Big{(}n-\frac{1}{n}\Big{)}=\tilde{h}^{\sigma}_{n}. (11)

The nn-twisted Ramond sector is generated by twisted spin operators with the appropriate 1n\frac{1}{n} rescaling of their weights. For the representative permutation (1n)(1\cdots n), the nn-twisted R-charged fields Rnα(z)R^{\alpha}_{n}(z) are

Rn±(z)\displaystyle R_{n}^{\pm}(z) exp(±i2nI=1n[ϕ1,I(z)ϕ2,I(z)])σ(1n)(z)\displaystyle\equiv\exp\left(\pm\frac{i}{2n}\sum_{I=1}^{n}\big{[}\phi_{1,I}(z)-\phi_{2,I}(z)\big{]}\right)\sigma_{(1\cdots n)}(z) (12)

with a similar construction for the neutral RnA˙(z)R^{\dot{A}}_{n}(z). From these, we can compose SNS_{N}-invariant combinations R[n]ζ(z)R^{\zeta}_{[n]}(z), ζ=±,A˙\zeta=\pm,\dot{A} by summing over orbits, as we did for the normalized SNS_{N}-invariant twists σ[n]\sigma_{[n]}. For example, the R-charged fields, with which we will be primarily concerned, are written explicitly as

R[n]±(z)1𝒮n(N)hSNexp(±i2nI=1n[ϕ1,h(I)(z)ϕ2,h(I)(z)])σh1(1n)h(z)R^{\pm}_{[n]}(z)\equiv\frac{1}{\mathscr{S}_{n}(N)}\sum_{h\in S_{N}}\exp\left(\pm\frac{i}{2n}\textstyle\sum_{I=1}^{n}\big{[}\phi_{1,h(I)}(z)-\phi_{2,h(I)}(z)\big{]}\right)\sigma_{h^{-1}(1\cdots n)h}(z) (13)

where in the exponential we sum over h(I)={h(1),,h(n)}h(I)=\{h(1),\cdots,h(n)\}, the image of the original copy set I={1,,n}I=\{1,\cdots,n\} under the permutation hh. The R[n]±R^{\pm}_{[n]}, like the spin fields S±S^{\pm}, form a doublet of SU(2)L and a singlet of SU(2)1, with charges j3=±12j^{3}=\pm\tfrac{1}{2} and 𝔧3=0\mathfrak{j}^{3}=0. On the other hand, the R[n]A˙(z)R^{\dot{A}}_{[n]}(z) form a singlet of R-symmetry and a doublet of SU(2)1, with charges j3=0j^{3}=0 and 𝔧3=±12\mathfrak{j}^{3}=\pm\frac{1}{2}. The conformal weight of the Rn(ζ)R^{(\zeta)}_{n} is

hnR=n4,h^{R}_{n}=\frac{n}{4}, (14)

obtained from the combined weights of the exponential and the twist. Completely analogous fields R~[n]ζ(z¯)\tilde{R}^{\zeta}_{[n]}(\bar{z}), with dimension h~nR=hnR\tilde{h}^{R}_{n}=h^{R}_{n}, make the anti-holomorphic sector. The normalization factor 𝒮n(N)\mathscr{S}_{n}(N) ensures that the two-point functions are normalized, granted that the non-SNS_{N}-invariant functions are normalized:

R[n]()R[n]±(0)=1=Rn()Rn±(0).\displaystyle\big{\langle}R^{\mp}_{[n]}(\infty)\,R^{\pm}_{[n]}(0)\big{\rangle}=1=\big{\langle}R^{\mp}_{n}(\infty)\,R^{\pm}_{n}(0)\big{\rangle}. (15)

Let us examine these fields a little further. The Hilbert space of the orbifold theory, orb=[g]g\mathcal{H}_{orb}=\oplus_{[g]}\mathcal{H}_{g}, is a direct sum of sectors invariant under the conjugacy classes of SNS_{N} Dijkgraaf:1996xw . The latter are given by the irreducible decomposition of gSNg\in S_{N} into disjoint cycles, [g]=k(k)qk[g]=\prod_{k\in\mathbb{N}}(k)^{q_{k}}, with kkqk=N\sum_{k}kq_{k}=N, and we are interested in the simplest sector, corresponding to [g]=(1)Nn(n)[g]=(1)^{N-n}(n). This Hilbert space, (n)\mathcal{H}_{(n)}, is invariant under the centralizer subgroup SNn×nS_{N-n}\times\mathbb{Z}_{n}, where SNnS_{N-n} permutes the NnN-n trivial cycles (1)(1), and n\mathbb{Z}_{n} acts on the elements permuted by (n)(n). It can be further decomposed as Dijkgraaf:1996xw

(n)=𝒮(Nn)(n)n,\mathcal{H}_{(n)}=\mathcal{S}^{(N-n)}\otimes\mathcal{H}^{\mathbb{Z}_{n}}_{(n)},

where 𝒮(Nn)\mathcal{S}^{(N-n)} is the symmetric tensor product of copies entering the trivial cycles, and where (n)n\mathcal{H}^{\mathbb{Z}_{n}}_{(n)} corresponds to the copies permuted by (n)n(n)\in\mathbb{Z}_{n}. States in (n)\mathcal{H}_{(n)} can be interpreted as a string with winding number nn. We can think of the construction of the operators Rn(ζ)R^{(\zeta)}_{n} as exciting the nn-wound copies to the Ramond ground state, while leaving the unwound copies in the NS vacuum. Thus the operators Rn(ζ)R^{(\zeta)}_{n} correspond to states |Rn(ζ)=|N1NS|,ζ(n)R\ket{R^{(\zeta)}_{n}}=\ket{\varnothing}^{\mathrm{NS}}_{N-1}\otimes\ket{\varnothing,\zeta}^{\mathrm{R}}_{(n)} where |N1NS𝒮N1\ket{\varnothing}^{\mathrm{NS}}_{N-1}\in\mathcal{S}^{N-1} is the NS vacuum of the non-twisted copies and |,ζ(n)R\ket{\varnothing,\zeta}^{\mathrm{R}}_{(n)} are Ramond ground states of a CFT defined on (n)n\mathcal{H}^{\mathbb{Z}_{n}}_{(n)}. Since the latter CFT involves nn copies of the 𝒩=4\mathcal{N}=4 SCFT, it has central charge c=6nc=6n; its Ramond ground states have the conformal weight c24=n4\frac{c}{24}=\frac{n}{4} in Eq.(14). We will sometimes refer to R[n]αR^{\alpha}_{[n]} as ‘Ramond ground states’, but it should be kept in mind that this is an abuse of nomenclature, as RnαR^{\alpha}_{n} are ground states only of the nn-wound string; the true Ramond ground states of the orbifold theory have conformal weight h=124corb=14Nh=\frac{1}{24}c_{orb}=\frac{1}{4}N, and are made either by R[N]αR^{\alpha}_{[N]}, i.e. by single-cycle fields with maximal twist n=Nn=N, or, more generally, by “composite fields” k(Rnkαk)k\prod_{k}(R^{\alpha_{k}}_{n_{k}})^{k} with kknk=N\sum_{k}kn_{k}=N.

The main objective of this paper is to describe how the dimensions hnR=14nh^{R}_{n}=\frac{1}{4}n are corrected when the free orbifolded SCFT is perturbed by a marginal operator.

3 Away from the free orbifold

A marginal deformation of the free orbifold turns the theory into an interacting SCFT, with the action

Sint=Sfree+λd2zO[2](int)(z,z¯)S_{\mathrm{int}}=S_{\mathrm{free}}+\lambda\int d^{2}z\,O^{(\mathrm{int})}_{[2]}(z,\bar{z}) (16)

parameterized by a dimensionless deformation parameter λ\lambda. In the large-NN limit, in which we will be interested, the deformation parameter λ\lambda should scale with NN in such a way that the ’t Hooft coupling λλ/N\lambda_{*}\equiv\lambda/\sqrt{N} is held fixed as NN\to\infty; see Lunin:2000yv ; Pakman:2009zz .

The “scalar modulus” interaction operator O[2](int)O^{(\mathrm{int})}_{[2]} is marginal, with total conformal dimension Δ=h+h~=2\Delta=h+\tilde{h}=2. This dimension should not change under renormalization. Also, O[2](int)O^{(\mathrm{int})}_{[2]} must be a singlet of R-symmetry, in order for 𝒩=(4,4)\mathcal{N}=(4,4) SUSY not to be broken. From the 20 deformation operators, which correspond to the 20 SUGRA moduli (see Avery:2010er ), we consider the SNS_{N}-invariant singlet

O[2](int)(z,z¯)=ϵABG12AG~12˙BO[2](0,0)(z,z¯)O^{(\mathrm{int})}_{[2]}(z,\bar{z})=\epsilon_{AB}G^{-A}_{-\frac{1}{2}}\tilde{G}^{\dot{-}B}_{-\frac{1}{2}}O^{(0,0)}_{[2]}(z,\bar{z}) (17)

constructed as a descendent of the NS chiral field O[2](0,0)(z)O^{(0,0)}_{[2]}(z) with h=12=j3h=\frac{1}{2}=j^{3}.


Let us review a few key results in conformal perturbation theory used in the next sections; see for example Keller:2019yrr for more detail. For a marginal perturbation, the two-point function

𝒪(z1,z¯1)𝒪(z2,z¯2)λ=|z12|2Δλ\big{\langle}\mathscr{O}(z_{1},\bar{z}_{1})\mathscr{O}(z_{2},\bar{z}_{2})\big{\rangle}_{\lambda}=|z_{12}|^{-2\Delta_{\lambda}} (18)

of a neutral and hermitian (for simplicity) operator 𝒪\mathscr{O} is still fixed by conformal symmetry, hence the effect of the marginal perturbation has to be a change of its conformal dimension. The λ\lambda expansion of the functional integral gives

𝒪(z1,z¯1)𝒪(z2,z¯2)λ=1|z12|2Δ[1+2πCλlog|z12|Λ+πλ2log|z12|Λd2uG(u,u¯)+O(λ3)],\begin{split}\big{\langle}\mathscr{O}(z_{1},\bar{z}_{1})\mathscr{O}(z_{2},\bar{z}_{2})\big{\rangle}_{\lambda}=\frac{1}{|z_{12}|^{2\Delta}}\Bigg{[}1&+2\pi C\lambda\log\frac{|z_{12}|}{\Lambda}\\ &+\pi\lambda^{2}\log\frac{|z_{12}|}{\Lambda}\int\!d^{2}u\,G(u,\bar{u})+\mathrm{O}(\lambda^{3})\Bigg{]},\end{split} (19)

where absence of a λ\lambda-index (e.g. in Δ\Delta) always indicates evaluation in the free theory, and the objects in the r.h.s. are defined as follows. At first order, CC is the structure constant coming from the three-point function

C=𝒪()O[2](int)(1)𝒪(0).C=\big{\langle}\mathscr{O}(\infty)O^{(\mathrm{int})}_{[2]}(1)\mathscr{O}(0)\big{\rangle}. (20)

At second order, G(u,u¯)G(u,\bar{u}) is the undetermined part of the four-point function in terms of the anharmonic ratio u(z12z34)/(z13z24)u\equiv(z_{12}z_{34})/(z_{13}z_{24}),

𝒪(z1,z¯1)O[2](int)(z3,z¯3)O[2](int)(z4,z¯4)𝒪(z2,z¯2)=G(u,u¯)|z13|2|z32|2|z12|2Δ2,\big{\langle}\mathscr{O}(z_{1},\bar{z}_{1})O^{(\mathrm{int})}_{[2]}(z_{3},\bar{z}_{3})O^{(\mathrm{int})}_{[2]}(z_{4},\bar{z}_{4})\mathscr{O}(z_{2},\bar{z}_{2})\big{\rangle}=\frac{G(u,\bar{u})}{|z_{13}|^{2}|z_{32}|^{2}|z_{12}|^{2\Delta-2}}, (21)

and Λ\Lambda is a cutoff for the integral

d2z3|z13|2|z32|2|z12|2Δ2=2π|z12|2Δlog|z12|Λ,Λ1.\int\!\frac{d^{2}z_{3}}{|z_{13}|^{2}|z_{32}|^{2}|z_{12}|^{2\Delta-2}}=\frac{2\pi}{|z_{12}|^{2\Delta}}\log\frac{|z_{12}|}{\Lambda},\qquad\quad\Lambda\ll 1. (22)

The logΛ\log\Lambda divergence in the two-point function requires the introduction of an appropriate regularization and a corresponding renormalization of the field 𝒪\mathscr{O}. The logarithmic form of the divergent terms indeed has the effect of changing the exponent of the renormalized two-point function, thus changing Δ\Delta. The operators we are interested in have a vanishing three-point function with O[2](int)O^{(\mathrm{int})}_{[2]}, i.e. C=0C=0. The corrections in (19) therefore start at second order in λ\lambda, and the renormalized field is

𝒪(ren)(z,z¯)=Λ12πλ2J𝒪(z,z¯),\displaystyle\mathscr{O}^{(ren)}(z,\bar{z})=\Lambda^{\frac{1}{2}\pi\lambda^{2}J}\mathscr{O}(z,\bar{z}), (23)
where Jd2uG(u,u¯).\displaystyle J\equiv\int\!d^{2}u\,G(u,\bar{u}). (24)

We can see that

𝒪(ren)(z1,z¯1)𝒪(ren)\displaystyle\big{\langle}\mathscr{O}^{(ren)}(z_{1},\bar{z}_{1})\mathscr{O}^{(ren)} (z2,z¯2)λ=Λπλ2J𝒪(z1,z¯1)𝒪(z2,z¯2)λ\displaystyle(z_{2},\bar{z}_{2})\big{\rangle}_{\lambda}=\Lambda^{\pi\lambda^{2}J}\big{\langle}\mathscr{O}(z_{1},\bar{z}_{1})\mathscr{O}(z_{2},\bar{z}_{2})\big{\rangle}_{\lambda}
=(1+πλ2JlogΛ+)|z12|2Δ(1+πλ2Jlog|z12|Λ+)\displaystyle=\Big{(}1+\pi\lambda^{2}J\log\Lambda+\cdots\Big{)}|z_{12}|^{-2\Delta}\Big{(}1+\pi\lambda^{2}J\log\frac{|z_{12}|}{\Lambda}+\cdots\Big{)}
=(1+πλ2Jlog|z12|+)|z12|2Δ\displaystyle=\Big{(}1+\pi\lambda^{2}J\log|z_{12}|+\cdots\Big{)}|z_{12}|^{-2\Delta}
=|z12|πλ2J|z12|2Δ\displaystyle=|z_{12}|^{\pi\lambda^{2}J}|z_{12}|^{-2\Delta}

so the Λ\Lambda-divergence is canceled, and the free-theory dimension Δ\Delta has flowed to a λ\lambda-dependent value

Δλ=Δπ2λ2J+O(λ3).\Delta_{\lambda}=\Delta-\tfrac{\pi}{2}\lambda^{2}J+\mathrm{O}(\lambda^{3}). (25)

The integral JJ also gives the first-order λ\lambda-correction to the particular structure constant in (20). This can be seen from the functional integral expansion of the corresponding three-point function. For our case where the free-theory constant vanishes, we find

Cλ=λJ+O(λ2).C_{\lambda}=\lambda J+\mathrm{O}(\lambda^{2}). (26)

To compute the integral (24), we need to be able to calculate the four-point function (21) in the free orbifold theory. In the next section, we show how to do this.

4 Four-point functions

Our goal is to compute four-point functions444A note on convention: in this paper, fields inside correlation functions are to be understood in two-dimensional theory, e.g. σn(z,z¯)\sigma_{n}(z,\bar{z}), instead of σn(z)\sigma_{n}(z). However, when fixing a point in 2{\mathbb{C}}^{2} we only write one argument for economy of notation. Thus, in (27), it should be understood that 𝒪[n](0)=𝒪[n](0,0¯)\mathscr{O}_{[n]}(0)=\mathscr{O}_{[n]}(0,\bar{0}), O[2](int)(1)=O[2](int)(1,1¯)O^{(\mathrm{int})}_{[2]}(1)=O^{(\mathrm{int})}_{[2]}(1,\bar{1}), etc.

G(u,u¯)=𝒪[n]()O[2](int)(1)O[2](int)(u,u¯)𝒪[n](0)G(u,\bar{u})=\big{\langle}\mathscr{O}_{[n]}^{\dagger}(\infty)O^{(\mathrm{int})}_{[2]}(1)O^{(\mathrm{int})}_{[2]}(u,\bar{u})\mathscr{O}_{[n]}(0)\big{\rangle} (27)

for primary operators 𝒪\mathscr{O} in the nn-twisted sectors of the orbifold SCFT2. Some aspects of the computation are universal, depending only on the nature of the twists: we start by describing the covering surface appropriate to the twisted structure of (27); then we describe the stress-tensor method to compute the simplest four-point function with this structure, containing only bare twists. Finally, we turn to the cases containing interaction operators with 𝒪[n]\mathscr{O}_{[n]} as the charged Ramond ground state, or as a bare twist field.

4.1 The covering surface

Twisted correlators such as (27) are complicated functions, with specific monodromies of their arguments fixed by their (bare) twist fields constituents. The standard way Lunin:2000yv of implementing the boundary conditions (8) for G(u,u¯)G(u,\bar{u}) is to map the ‘base sphere’ Sbase2=S^{2}_{\mathrm{base}}=\mathbb{C}\cup\infty to a ramified ‘covering surface’ Σcover\Sigma_{\mathrm{cover}}, whose ramification points correspond to the position and the order of twists operators. At large NN, the leading contribution comes from genus-zero covering surfaces. Denote coordinates on the base by zSbase2z\in S^{2}_{\mathrm{base}}, and coordinates on the covering sphere by tScover2t\in S^{2}_{\mathrm{cover}}, and fix the four punctures on each surface to be

{z=0}{t=0},{z=1}{t=t1},{z=u}{t=x},{z=}{t=}.\{z=0\}\mapsto\{t=0\},\{z=1\}\mapsto\{t=t_{1}\},\{z=u\}\mapsto\{t=x\},\{z=\infty\}\mapsto\{t=\infty\}.

The method for finding z(t)z(t) for generic monodromies was pioneered in Lunin:2000yv and generalized in Pakman:2009zz . For the specific monodromies (and topology) above,

z(t)=(tt1)n(tt0t1t0)(t1ttt).z(t)=\left({t\over t_{1}}\right)^{n}\left(\frac{t-t_{0}}{t_{1}-t_{0}}\right)\left(\frac{t_{1}-t_{\infty}}{t-t_{\infty}}\right). (28)

The monodromies at z=0z=0 and z=z=\infty are evident, but at z=1z=1 and z=uz=u they are implicit in the derivative z(t)z^{\prime}(t), which must vanish at every branching point. Indeed,

dzdt=t1tt1(t0t1)tn1(tt)2[(tt0)tn(tt0)(tt)]\frac{dz}{dt}=\frac{t_{1}-t_{\infty}}{t_{1}(t_{0}-t_{1})}\frac{t^{n-1}}{(t-t_{\infty})^{2}}\Big{[}(t_{\infty}-t_{0})t-n(t-t_{0})(t-t_{\infty})\Big{]} (29)

vanishes at t=0t=0 with the correct monodromy, while xx and t1t_{1} must be the roots of the quadratic expression in brackets. This quadratic equation relates the parameters t1,t0,tt_{1},t_{0},t_{\infty} and xx,

x+t1=n1nt0+n+1nt;xt1=t0t.x+t_{1}=\tfrac{n-1}{n}t_{0}+\tfrac{n+1}{n}t_{\infty};\qquad xt_{1}=t_{0}t_{\infty}. (30)

We are free to choose one of the ratios t0/xt_{0}/x, t0/t1t_{0}/t_{1}, t/t1t_{\infty}/t_{1} and t/xt_{\infty}/x as long as they satisfy the two conditions (30), and we choose

1t0x=1x,hence1tx=1x+n,1t1x=2x+n1x(n+x)1-\frac{t_{0}}{x}=\frac{1}{x},\quad\text{hence}\quad 1-\frac{t_{\infty}}{x}=\frac{1}{x+n},\quad 1-\frac{t_{1}}{x}=\frac{2x+n-1}{x(n+x)} (31)

which gives u=z(x)u=z(x) as Arutyunov:1997gi ; Arutyunov:1997gt

u(x)=xn1(x+n)n+1(x1)n+1(x+n1)n1,u(x)={x^{n-1}(x+n)^{n+1}\over(x-1)^{n+1}(x+n-1)^{n-1}}, (32)

a rational function.

4.2 Four-point functions

The covering map encodes the monodromies of functions like (27), with the twist structure

g(u,u¯)σn()σ2(1)σ2(u,u¯)σn(0),g(u,\bar{u})\equiv\big{\langle}\sigma_{n}(\infty)\sigma_{2}(1)\sigma_{2}(u,\bar{u})\sigma_{n}(0)\big{\rangle}, (33)

into the ramification points of the covering surface. One way of computing g(u,u¯)g(u,\bar{u}), formulated by Lunin and Mathur Lunin:2000yv , is to cut circles around the ramification points, replace them with vacua and compute the functional integral directly. An alternative555Still other ways of computing general four-point functions σmσnσpσq\langle\sigma_{m}\sigma_{n}\sigma_{p}\sigma_{q}\rangle have been recently given Roumpedakis:2018tdb ; Dei:2019iym . Dixon:1986qv is to use the conformal Ward identity: if one is able to find the residue r(u)r(u) of the following function on the base,

f(z)=T(z)σn()σ2(1)σ2(u,u¯)σn(0)σn()σ2(1)σ2(u,u¯)σn(0)=h(zu)2+r(u)zu+non-sing.,\begin{split}f(z)&=\frac{\big{\langle}T(z)\sigma_{n}(\infty)\sigma_{2}(1)\sigma_{2}(u,\bar{u})\sigma_{n}(0)\big{\rangle}}{\big{\langle}\sigma_{n}(\infty)\sigma_{2}(1)\sigma_{2}(u,\bar{u})\sigma_{n}(0)\big{\rangle}}=\frac{h}{(z-u)^{2}}+\frac{r(u)}{z-u}+\text{non-sing.},\end{split} (34)

the Ward identity gives a differential equation

ulogg(u)=r(u)\partial_{u}\log g(u)=r(u) (35)

which can be solved for the holomorphic part of g(u,u¯)=g(u)g~(u¯)g(u,\bar{u})=g(u)\tilde{g}(\bar{u}). The anti-holomorphic part g~(u¯)=g¯(u¯)\tilde{g}(\bar{u})=\bar{g}(\bar{u}) is obtained likewise, using T~(z¯)\tilde{T}(\bar{z}).

In simpler orbifold theories, it is possible to find r(u)r(u) by engineering the function with the appropriate poles and monodromies Dixon:1986qv . Here, we can follow Refs.Arutyunov:1997gi ; Arutyunov:1997gt ; Pakman:2009ab ; Pakman:2009zz and use the covering surface as an aid, by computing the correlation functions on Scover2S^{2}_{\mathrm{cover}}, where the monodromies are trivial, and then mapping back: f(z)f(z) is a function of the position of the stress tensor which, unlike the twists, is not placed on a branching/ramification point — hence mapping from covering to base is just a conformal transformation. On the covering,

fcover(t)=T(t)𝟙𝟙=0f_{\mathrm{cover}}(t)=\frac{\langle T(t)\mathds{1}\rangle}{\langle\mathds{1}\rangle}=0 (36)

because the twists disappear, and when mapping back to base only the anomalous transformation of TT does not cancel in the fraction, so

f(z)=I[c12{tI,z}+(dtIdz)2fcover(tI(z))]=I12{tI,z}.f(z)=\sum_{I}\left[\frac{c}{12}\{t_{I},z\}+\left(\frac{dt_{I}}{dz}\right)^{2}f_{\mathrm{cover}}(t_{I}(z))\right]=\sum_{I}\tfrac{1}{2}\{t_{I},z\}. (37)

The position of the twists appear as parameters implicit in the inverse maps ztz\mapsto t, which encode the twist structure of (33). There is a sum over II in Eq.(37) because T(z)T(z) is a sum over copies (7). Around a branching point, there is one inverse map tI(z)t_{I}(z) for each copy entering the corresponding twist; at z=uz=u, the insertion point of σ2\sigma_{2}, there are two maps, which can be found locally Arutyunov:1997gt ; Pakman:2009mi , as follows. Take the logarithm of the ratio z(t)/z(x)z(t)/z(x), i.e.  log(z/u)=nlogtx+logtt0xt0logttxt\log(z/u)=n\log\frac{t}{x}+\log\frac{t-t_{0}}{x-t_{0}}-\log\frac{t-t_{\infty}}{x-t_{\infty}}, and expand both sides,

k=1bk(zu)k=(tx)2k=0ak(tx)k,hencetx=k=1ck(zu)k/2.\sum_{k=1}^{\infty}b_{k}(z-u)^{k}=(t-x)^{2}\sum_{k=0}^{\infty}a_{k}(t-x)^{k},\quad\text{hence}\quad t-x=\sum_{k=1}^{\infty}c_{k}(z-u)^{k/2}. (38)

In the first equation, the coefficients are found from the Taylor expansions,

bk=(1)k+1kuk,ak=(1)k+1k+2[1(xt0)k+21(xt)k+2+nxk+2].\displaystyle b_{k}=\frac{(-1)^{k+1}}{ku^{k}},\quad a_{k}=\frac{(-1)^{k+1}}{k+2}\left[\frac{1}{(x-t_{0})^{k+2}}-\frac{1}{(x-t_{\infty})^{k+2}}+\frac{n}{x^{k+2}}\right]. (39)

The coefficients ckc_{k} are solved in terms of aka_{k} and bkb_{k} order by order, by inserting the ckc_{k} power series into the first equation in (38). The multiple inverses ztz\mapsto t appear as multiple solutions for the ckc_{k}. After solving for the ckc_{k}, we can put the powers series into the r.h.s. of Eq.(37), expand to order (zu)1(z-u)^{-1} and extract the desired residue. The coefficients c1c_{1}, c2c_{2} and c3c_{3} completely determine the result up to this order,

c1=±b1a0,c2=a1b12a02,c3=±4a0a2b124a03b25a12b128a07/2b1.\displaystyle c_{1}=\pm\sqrt{\frac{b_{1}}{a_{0}}},\quad c_{2}=-\frac{a_{1}b_{1}}{2a_{0}^{2}},\quad c_{3}=\pm\frac{4a_{0}a_{2}b_{1}^{2}-4a_{0}^{3}b_{2}-5a_{1}^{2}b_{1}^{2}}{8a_{0}^{7/2}\sqrt{b_{1}}}. (40)

As expected, there are two solutions. When the parameters t0t_{0} and tt_{\infty} in ak,bka_{k},b_{k} are written explicitly in terms of xx, these coefficients are functions of xx alone, thus we find the residue rr as a function of xx. One can check that r(x)r(x) is the same for both choices of the ckc_{k}.

Solving Eq.(35) requires expressing r(x)r(x) as an explicit function of uu, but there are multiple inverses of u(x)u(x). It is easier to make a change of variables, and solve instead the differential equation

xlogg(x)=u(x)r(x),\partial_{x}\log g(x)=u^{\prime}(x)r(x), (41)

whose solution is

g(x)=cσx2+5n(n1)8n(x1)2+5n(n+1)8n(x+n)2n(n+1)8n(x+n1)2n(n1)8n(x+n12)1/4.g(x)=c_{\sigma}\frac{x^{-\frac{2+5n(n-1)}{8n}}(x-1)^{\frac{2+5n(n+1)}{8n}}(x+n)^{\frac{2-n(n+1)}{8n}}(x+n-1)^{-\frac{2-n(n-1)}{8n}}}{(x+\frac{n-1}{2})^{1/4}}. (42)

The integration constant cσc_{\sigma} has to be determined by looking at OPE limits (see App.C).666We emphasize that the function σn()σ2(1)σ2(u,u¯)σn(0)\langle\sigma_{n}(\infty)\sigma_{2}(1)\sigma_{2}(u,\bar{u})\sigma_{n}(0)\rangle is known in the literature, calculated by other methods. Our point is to take it as an instructive example of the specific method we use.

Now, we have found a function parameterized by the pre-image of uu under the covering map z(t)z(t). For fixed u=uu=u_{*}, there are 𝐇\bf H different pre-images x𝔞x_{\mathfrak{a}}, 𝔞=1,,𝐇\mathfrak{a}=\mathfrak{1},\cdots,\bf H, solutions of the equation

xn1(x+n)n+1u(x1)n+1(x+n1)n1=0.x^{n-1}(x+n)^{n+1}-u_{*}(x-1)^{n+1}(x+n-1)^{n-1}=0. (43)

The degree of the polynomial shows that 𝐇=2n{\bf H}=2n. Note that this is not the number of sheets of the ramified covering (uu is the position of a branching point), it is the number of different covering maps with the assumed monodromy conditions; 𝐇\bf H is a Hurwitz number Pakman:2009zz ; Pakman:2009ab ; Pakman:2009mi .

The method has thus yielded 𝐇\bf H functions g(x𝔞(u))g(x_{\mathfrak{a}}(u)). This was expected, because the SNS_{N} structure of the composition of cycles in Eq.(33) is not completely fixed. Labeling cycles by the position of their twists operators, those entering g(u,u¯)g(u,\bar{u}) must compose to the identity,

(n)(2)1(2)u(n)0=1,(n)_{\infty}(2)_{1}(2)_{u}(n)_{0}=1, (44)

otherwise the correlator vanishes. There are several collections {(n),(2)1,(2)u,(n)0}\{(n)_{\infty},(2)_{1},(2)_{u},(n)_{0}\} of cycles which solve Eq.(44),777The total number of such solutions can be found with Frobenius’ formula lando2013graphs . and these collections can be arranged into equivalence classes defined by

(n)(2)1(2)u(n)0h(n)h1h(2)1h1h(2)uh1h(n)0h1hSN.(n)_{\infty}(2)_{1}(2)_{u}(n)_{0}\sim h(n)_{\infty}h^{-1}\,h(2)_{1}h^{-1}\ h(2)_{u}h^{-1}\,h(n)_{0}h^{-1}\quad\forall\quad h\in S_{N}. (45)

The existence of different such equivalence classes is the reason for the existence of different functions g(x𝔞(u))g(x_{\mathfrak{a}}(u)); there are precisely 𝐇=2n{\bf H}=2n equivalence classes Pakman:2009zz . Inside each of these classes, let C𝐬(N)C_{\bf s}(N) be the number of collections {(n),(2)1,(2)u,(n)0}\{(n)_{\infty},(2)_{1},(2)_{u},(n)_{0}\} for which the cycles involve a fixed number 𝐬{\bf s} of distinct elements of {1,2,,N}\{1,2,\cdots,N\}. Then it can be shown Pakman:2009zz that C𝐬C_{\bf s} is the same for all classes, and that, for large NN, it scales as

C𝐬=N𝐬12r=14nr[ϖ(nr)+O(1/N)],C_{\bf s}=N^{{\bf s}-\frac{1}{2}\sum_{r=1}^{4}n_{r}}\left[\varpi(n_{r})+\mathrm{O}(1/N)\right], (46)

where n1=n=n4n_{1}=n=n_{4} and n2=2=n3n_{2}=2=n_{3} are the order of the q=4q=4 twists involved in (33). But nr1n_{r}-1 is also the order of the ramification points of the covering surface, 𝐬\bf s is the number of its sheets, hence its genus is fixed by the Riemann-Hurwitz formula

𝐠=1𝐬+12r=1q(nr1).{\bf g}=1-{\bf s}+\frac{1}{2}\sum_{r=1}^{q}(n_{r}-1). (47)

We thus see that C𝐬(N)C𝐠(N)N𝐠1,C_{{\bf s}}(N)\equiv C_{{\bf g}}(N)\sim N^{-{\bf g}-1}, therefore the covering surface with 𝐠=0{\bf g}=0 constructed in §4.1 gives the leading contribution at large NN Lunin:2000yv . For our four-point functions, the Riemann-Hurwitz formula gives 𝐬=𝐠+n+1{\bf s}=-{\bf g}+n+1, hence we see that, for the covering surface to have genus zero, we must have

1n<N.1\leq n<N. (48)

When we sum over the orbits of individual cycles to make an SNS_{N}-invariant correlation function, we get all terms in each of the equivalence classes above,

σ[n]()σ[2](1)σ[2](u,u¯)σ[n](0)=ϖ(n)N𝔞=1𝐇g(x𝔞(u))g¯(x¯𝔞(u¯)).\big{\langle}\sigma_{[n]}(\infty)\sigma_{[2]}(1)\sigma_{[2]}(u,\bar{u})\sigma_{[n]}(0)\big{\rangle}=\frac{\varpi(n)}{N}\sum_{\mathfrak{a}=\mathfrak{1}}^{\bf H}g(x_{\mathfrak{a}}(u))\bar{g}(\bar{x}_{\mathfrak{a}}(\bar{u})). (49)

This sum corresponds to different OPE channels resulting from composing the twist permutations, not only for g(u,u¯)g(u,\bar{u}) but for the other functions G(u,u¯)G(u,\bar{u}) which share the same twist structure.

4.2.1 Charged Ramond fields

Let us now turn to the function

GR(u,u¯)=R[n]()O[2](int)(1)O[2](int)(u,u¯)R[n]+(0).G_{R}(u,\bar{u})=\big{\langle}R^{-}_{[n]}(\infty)O^{(\mathrm{int})}_{[2]}(1)O^{(\mathrm{int})}_{[2]}(u,\bar{u})R^{+}_{[n]}(0)\big{\rangle}. (50)

The Ramond fields R[n]±(z,z¯)R^{\pm}_{[n]}(z,\bar{z}) are lifted to the corresponding spin field S±(t,t¯)S^{\pm}(t,\bar{t}), so we compute

Fcover(t)=T(t)S()O(int)(t1,t¯1)O(int)(x,x¯)S+(0)S()O(int)(t1,t¯1)O(int)(x,x¯)S+(0)F_{\mathrm{cover}}(t)=\frac{\big{\langle}T(t)S^{-}(\infty)O^{(\mathrm{int})}(t_{1},\bar{t}_{1})O^{(\mathrm{int})}(x,\bar{x})S^{+}(0)\big{\rangle}}{\big{\langle}S^{-}(\infty)O^{(\mathrm{int})}(t_{1},\bar{t}_{1})O^{(\mathrm{int})}(x,\bar{x})S^{+}(0)\big{\rangle}} (51)

and then find the residue HH of the function

F(z)=T(z)R[n]()O[2](int)(1)O[2](int)(u,u¯)R[n]+(0)R[n]()O[2](int)(1)O[2](int)(u,u¯)R[n]+(0)=2[12{t,z}+(dtdz)2Fcover(t(z),x)]=H(x)zu+\begin{split}F(z)&=\frac{\big{\langle}T(z)R^{-}_{[n]}(\infty)O^{(\mathrm{int})}_{[2]}(1)O^{(\mathrm{int})}_{[2]}(u,\bar{u})R^{+}_{[n]}(0)\big{\rangle}}{\big{\langle}R^{-}_{[n]}(\infty)O^{(\mathrm{int})}_{[2]}(1)O^{(\mathrm{int})}_{[2]}(u,\bar{u})R^{+}_{[n]}(0)\big{\rangle}}\\ &=2\left[\frac{1}{2}\big{\{}t,z\big{\}}+\left(\frac{dt}{dz}\right)^{2}F_{\mathrm{cover}}(t(z),x)\right]=\frac{H(x)}{z-u}+\cdots\end{split} (52)

with t(z)t(z) one of the maps obtained from Eqs.(38) and (40).

The deformation operator, denoted by O(int)(t,t¯)O^{(\mathrm{int})}(t,\bar{t}) — without a twist index since there are no twists on the covering surface — can be expressed on Scover2S^{2}_{\mathrm{cover}} in terms of the basic fields only, because the contour integrals in the super-current modes G12αA=12πi𝑑zGαA(z)G^{\alpha A}_{-\frac{1}{2}}=\frac{1}{2\pi i}\oint dzG^{\alpha A}(z) just pick up a residue (see e.g. Burrington:2012yq ). The result is a sum of products of bosonic currents, free fermions and spin fields coming from the lifting of the NS chiral field O[2](0,0)(z,z¯)S+(t)S~+˙(t¯)O^{(0,0)}_{[2]}(z,\bar{z})\mapsto S^{+}(t)\tilde{S}^{\dot{+}}(\bar{t}). Writing spin fields as exponentials,

O(int)=aint[:X1˙1e+i2(ϕ1+ϕ2)(¯X1˙2e+i2(ϕ~1+ϕ~2)(¯X1˙1)ei2(ϕ~1+ϕ~2))::X1˙2e+i2(ϕ1+ϕ2)((¯X1˙2)ei2(ϕ~1+ϕ~2)+¯X1˙1e+i2(ϕ~1+ϕ~2)):+:(X1˙1)ei2(ϕ1+ϕ2)((¯X1˙2)ei2(ϕ~1+ϕ~2)+¯X1˙1e+i2(ϕ~1+ϕ~2)):+:(X1˙2)ei2(ϕ1+ϕ2)(¯X1˙2e+i2(ϕ~1+ϕ~2)(¯X1˙1)ei2(ϕ~1+ϕ~2)):].\begin{split}O^{(\mathrm{int})}=a_{\mathrm{int}}\Big{[}&\,\colon\!\partial X^{\dot{1}1}\,e^{+\frac{i}{2}(\phi_{1}+\phi_{2})}\left(\bar{\partial}X^{\dot{1}2}e^{+\frac{i}{2}(\tilde{\phi}_{1}+\tilde{\phi}_{2})}-(\bar{\partial}X^{\dot{1}1})^{\dagger}e^{-\frac{i}{2}(\tilde{\phi}_{1}+\tilde{\phi}_{2})}\right)\!\colon\,\\ -&\,\colon\!\partial X^{\dot{1}2}\,e^{+\frac{i}{2}(\phi_{1}+\phi_{2})}\left((\bar{\partial}X^{\dot{1}2})^{\dagger}e^{-\frac{i}{2}(\tilde{\phi}_{1}+\tilde{\phi}_{2})}+\bar{\partial}X^{\dot{1}1}e^{+\frac{i}{2}(\tilde{\phi}_{1}+\tilde{\phi}_{2})}\right)\!\colon\,\\ +&\,\colon\!(\partial X^{\dot{1}1})^{\dagger}e^{-\frac{i}{2}(\phi_{1}+\phi_{2})}\left((\bar{\partial}X^{\dot{1}2})^{\dagger}e^{-\frac{i}{2}(\tilde{\phi}_{1}+\tilde{\phi}_{2})}+\bar{\partial}X^{\dot{1}1}e^{+\frac{i}{2}(\tilde{\phi}_{1}+\tilde{\phi}_{2})}\right)\!\colon\,\\ +&\,\colon\!(\partial X^{\dot{1}2})^{\dagger}\,e^{-\frac{i}{2}(\phi_{1}+\phi_{2})}\left(\bar{\partial}X^{\dot{1}2}e^{+\frac{i}{2}(\tilde{\phi}_{1}+\tilde{\phi}_{2})}-(\bar{\partial}X^{\dot{1}1})^{\dagger}e^{-\frac{i}{2}(\tilde{\phi}_{1}+\tilde{\phi}_{2})}\right)\!\colon\,\Big{]}.\end{split} (53)

The constant ainta_{\mathrm{int}} can be conveniently chosen by a redefinition of the deformation parameter λ\lambda. For now, we leave it unspecified. To compute the correlators, the strategy is to show that contractions of T(t)T(t) with the fields in the numerator of (51) are always proportional to 𝒢=S()O(int)(t1,t¯1)O(int)(x,x¯)S+(0)\mathcal{G}=\langle S^{-}(\infty)O^{(\mathrm{int})}(t_{1},\bar{t}_{1})O^{(\mathrm{int})}(x,\bar{x})S^{+}(0)\rangle, appearing in the denominator of Eq.(51). We can decompose T(t)=TB(t)+TF(t)T(t)=T_{B}(t)+T_{F}(t) into bosonic and fermionic parts, respectively

TB(t)\displaystyle T_{B}(t) =14ϵA˙B˙ϵAB:XA˙A(t)XB˙B(t):\displaystyle=-\tfrac{1}{4}\epsilon_{\dot{A}\dot{B}}\epsilon_{AB}\colon\!\partial X^{\dot{A}A}(t)\partial X^{\dot{B}B}(t)\colon
TF(t)\displaystyle T_{F}(t) =12:[ϕ1(t)ϕ1(t)+ϕ2(t)ϕ2(t)]:\displaystyle=-\tfrac{1}{2}\colon\!\!\big{[}\partial\phi_{1}(t)\partial\phi_{1}(t)+\partial\phi_{2}(t)\partial\phi_{2}(t)\big{]}\!\colon

As far as bosons are concerned, each term of the product O(int)(t1)O(int)(x)O^{(\mathrm{int})}(t_{1})O^{(\mathrm{int})}(x) has the structure XC˙C(t1)XE˙E(x)\partial X^{\dot{C}C}(t_{1})\partial X^{\dot{E}E}(x) multiplied by “transparent” fermionic or anti-holomorphic factors. Using the conformal Ward identity and the two-point functions (160),

TB(t)XC˙C(t1)XE˙E(x)\displaystyle\big{\langle}T_{B}(t)\partial X^{\dot{C}C}(t_{1})\partial X^{\dot{E}E}(x)\big{\rangle} =[t1tt1+xtx+1(tt1)2+1(tx)2]2ϵC˙E˙ϵCE(t1x)2\displaystyle=\left[\frac{\partial_{t_{1}}}{t-t_{1}}+\frac{\partial_{x}}{t-x}+\frac{1}{(t-t_{1})^{2}}+\frac{1}{(t-x)^{2}}\right]\frac{2\epsilon^{\dot{C}\dot{E}}\epsilon^{CE}}{(t_{1}-x)^{2}}
=(t1x)2(tt1)2(tx)2XC˙C(t1)XE˙E(x).\displaystyle=\frac{(t_{1}-x)^{2}}{(t-t_{1})^{2}(t-x)^{2}}\ \big{\langle}\partial X^{\dot{C}C}(t_{1})\partial X^{\dot{E}E}(x)\big{\rangle}.

Hence we can recompose 𝒢\mathcal{G}, and obtain

TB(t)S()O(int)(t1,t¯1)O(int)(x,x¯)S+(0)S()O(int)(t1,t¯1)O(int)(x,x¯)S+(0)=(t1x)2(tt1)2(tx)2.\frac{\big{\langle}T_{B}(t)S^{-}(\infty)O^{(\mathrm{int})}(t_{1},\bar{t}_{1})O^{(\mathrm{int})}(x,\bar{x})S^{+}(0)\big{\rangle}}{\big{\langle}S^{-}(\infty)O^{(\mathrm{int})}(t_{1},\bar{t}_{1})O^{(\mathrm{int})}(x,\bar{x})S^{+}(0)\big{\rangle}}=\frac{(t_{1}-x)^{2}}{(t-t_{1})^{2}(t-x)^{2}}. (54)

For the fermionic part of the calculation, it is very helpful to organize O(int)O^{(\mathrm{int})} as

O(int)(t,t¯)\displaystyle O^{(\mathrm{int})}(t,\bar{t}) V(t,t¯)+V+(t,t¯),\displaystyle\equiv V_{-}(t,\bar{t})+V_{+}(t,\bar{t}), (55a)
where V+(t,t¯)\displaystyle V_{+}(t,\bar{t}) =[(a.h.)1˙1X1˙1(a.h.)1˙2X1˙2]:e+i2(ϕ1+ϕ2):\displaystyle=\Big{[}\big{(}a.h.)_{\dot{1}1}\partial X^{\dot{1}1}-\big{(}a.h.)_{\dot{1}2}\partial X^{\dot{1}2}\Big{]}\,\colon\!e^{+\frac{i}{2}(\phi_{1}+\phi_{2})}\!\colon\, (55b)
V(t,t¯)\displaystyle V_{-}(t,\bar{t}) =[(a.h.)1˙1(X1˙1)+(a.h.)1˙2(X1˙2)]:ei2(ϕ1+ϕ2):\displaystyle=\Big{[}\big{(}{a.h.})_{\dot{1}1\dagger}(\partial X^{\dot{1}1})^{\dagger}+\big{(}{a.h.})_{\dot{1}2\dagger}(\partial X^{\dot{1}2})^{\dagger}\Big{]}\,\colon\!e^{-\frac{i}{2}(\phi_{1}+\phi_{2})}\!\colon\, (55c)

the (a.h)(a.h)s being combinations of anti-holomorphic fields which can be read from (53). This makes it is clear that contractions with O[2](int)O^{(\mathrm{int})}_{[2]} are very simple, and

rlimvtϕr(v)ϕr(t)S()O(int)(t1)O(int)(x,x¯)S+(0)rlim\contractionϕ(t)rϕϕr(t)ϕr(v\displaystyle\textstyle\sum_{r}\lim_{v\to t}\big{\langle}\partial\phi_{r}(v)\partial\phi_{r}(t)S^{-}(\infty)O^{(\mathrm{int})}(t_{1})O^{(\mathrm{int})}(x,\bar{x})S^{+}(0)\big{\rangle}-\textstyle\sum_{r}\lim\contraction{}{\phi}{{}_{r}(t)}{\phi}\phi_{r}(t)\phi_{r}(v
=[(i/2)2t2+(i/2)2(tt1)2+(i/2)2(tx)2]𝒢\displaystyle=\Bigg{[}\frac{(i/2)^{2}}{t^{2}}+\frac{(i/2)^{2}}{(t-t_{1})^{2}}+\frac{(i/2)^{2}}{(t-x)^{2}}\Bigg{]}\mathcal{G}
+2(i/2)2(tt1)(tx)S()[V(t1)V+(t1)][V(x)V+(x)]S+(0)\displaystyle\quad+\frac{2(i/2)^{2}}{(t-t_{1})(t-x)}\big{\langle}S^{-}(\infty)\big{[}V_{-}(t_{1})-V_{+}(t_{1})\big{]}[V_{-}(x)-V_{+}(x)\big{]}S^{+}(0)\big{\rangle} (56)

The second line in the r.h.s. can be further simplified because, since the only non-vanishing two-point functions (162) are between a field and its conjugate, it follows that V±(t,t¯)V±(v,v¯)=0\langle V_{\pm}(t,\bar{t})V_{\pm}(v,\bar{v})\rangle=0, hence

S()[V(t1)V+(t1)][V(x)V+(x)]S+(0)=𝒢.\big{\langle}S^{-}(\infty)\big{[}V_{-}(t_{1})-V_{+}(t_{1})\big{]}[V_{-}(x)-V_{+}(x)\big{]}S^{+}(0)\big{\rangle}=-\mathcal{G}. (57)

Putting this back in (56), 𝒢\mathcal{G} appears as a common factor canceled in (51),

TF(t)S()O(int)(t1,t¯1)O(int)(x,x¯)S+(0)S()O(int)(t1,t¯1)O(int)(x,x¯)S+(0)=14[1t2+(1tt11tx)2].\begin{split}&\frac{\big{\langle}T_{F}(t)S^{-}(\infty)O^{(\mathrm{int})}(t_{1},\bar{t}_{1})O^{(\mathrm{int})}(x,\bar{x})S^{+}(0)\big{\rangle}}{\big{\langle}S^{-}(\infty)O^{(\mathrm{int})}(t_{1},\bar{t}_{1})O^{(\mathrm{int})}(x,\bar{x})S^{+}(0)\big{\rangle}}=\frac{1}{4}\left[\frac{1}{t^{2}}+\left(\frac{1}{t-t_{1}}-\frac{1}{t-x}\right)^{2}\right].\end{split} (58)

Combining (54) and (58), we get

Fcover(t)=(t1x)2(tt1)2(tx)2+14[1t2+(1tt11tx)2].\begin{split}F_{\mathrm{cover}}(t)&=\frac{(t_{1}-x)^{2}}{(t-t_{1})^{2}(t-x)^{2}}+\frac{1}{4}\left[\frac{1}{t^{2}}+\left(\frac{1}{t-t_{1}}-\frac{1}{t-x}\right)^{2}\right].\end{split} (59)

Inverting the maps, we find the residue H(x)H(x) of F(z)F(z) to be

H(x)=[16x4+32(2n1)x3+4(2n1)(10n7)x2+4(n1)[10(n1)n+3]x+5(n2)(n1)2n][4n(n+2x1)3]1.\begin{split}H(x)=-\Big{[}&16x^{4}+32(2n-1)x^{3}\\ &+4(2n-1)(10n-7)x^{2}\\ &+4(n-1)\left[10(n-1)n+3\right]x\\ &+5(n-2)(n-1)^{2}n\Big{]}\big{[}4n(n+2x-1)^{3}\big{]}^{-1}.\end{split} (60)

The solution of the differential equation xlogGR(x)=u(x)H(x)\partial_{x}\log G_{R}(x)=u^{\prime}(x)H(x) is now easily found,

GR(x)=CRx5(2n)4(x1)5(2+n)4(x+n)23n4(x+n1)2+3n4(x+n12)4.G_{R}(x)=C_{R}\;\frac{x^{\frac{5(2-n)}{4}}(x-1)^{\frac{5(2+n)}{4}}(x+n)^{\frac{2-3n}{4}}(x+n-1)^{\frac{2+3n}{4}}}{(x+\frac{n-1}{2})^{4}}. (61)

where CRC_{R} is an integration constant.

4.2.2 Bare twists

Let us also consider

Gσ(u,u¯)=σ[n]()O[2](int)(1)O[2](int)(u,u¯)σ[n](0)G_{\sigma}(u,\bar{u})=\big{\langle}\sigma_{[n]}(\infty)O^{(\mathrm{int})}_{[2]}(1)O^{(\mathrm{int})}_{[2]}(u,\bar{u})\sigma_{[n]}(0)\big{\rangle} (62)

appearing in the second-order correction of the two-point function of bare twist fields. The computation of

Fcover(t)=T(t)𝟙()O(int)(t1,t¯1)O(int)(x,x¯)𝟙(0)𝟙()O(int)(t1,t¯1)O(int)(x,x¯)𝟙(0)=(t1x)2(tt1)2(tx)2+14(1tt11tx)2,\begin{split}F_{\mathrm{cover}}(t)&=\frac{\big{\langle}T(t)\mathds{1}(\infty)O^{(\mathrm{int})}(t_{1},\bar{t}_{1})O^{(\mathrm{int})}(x,\bar{x})\mathds{1}(0)\big{\rangle}}{\big{\langle}\mathds{1}(\infty)O^{(\mathrm{int})}(t_{1},\bar{t}_{1})O^{(\mathrm{int})}(x,\bar{x})\mathds{1}(0)\big{\rangle}}\\ &=\frac{(t_{1}-x)^{2}}{(t-t_{1})^{2}(t-x)^{2}}+\frac{1}{4}\left(\frac{1}{t-t_{1}}-\frac{1}{t-x}\right)^{2},\end{split} (63)

goes as before (but is simpler), and we find

Gσ(x)=Cσx110n+5n24n(x1)1+10n+5n24n(x+n)1+2n3n24n(x+n1)12n3n24n(x+n12)4.G_{\sigma}(x)=C_{\sigma}\;\frac{x^{-\frac{1-10n+5n^{2}}{4n}}(x-1)^{\frac{1+10n+5n^{2}}{4n}}(x+n)^{\frac{1+2n-3n^{2}}{4n}}(x+n-1)^{-\frac{1-2n-3n^{2}}{4n}}}{(x+\frac{n-1}{2})^{4}}. (64)

where CσC_{\sigma} is an integration constant. The same function has been computed in App.E of Ref.Pakman:2009mi , but using a different parameterization map u(x)u(x), in place of (32) (hence their function G(x)G(x) is different from ours).

5 OPE limits, fusion rules and structure constants

The short-distance behavior of G(u,u¯)G(u,\bar{u}) in the limits u1,0,u\to 1,0,\infty contains the complete conformal data of the operator product expansions of the fields involved — i.e. the OPE fusion rules. Recall that super-conformal invariance fixes the form of the OPE algebra of generic primary holomorphic fields 𝒪k(jk3)(u)\mathscr{O}^{(j^{3}_{k})}_{k}(u) with dimensions Δk\Delta_{k} and R-charges jk3j^{3}_{k} to be

𝒪1(j13)(u,u¯)𝒪2(j23)(0)=kC12k|u|ΔkΔ1Δ2𝒪k(jk3)(0)+descendants,\mathscr{O}_{1}^{(j^{3}_{1})}(u,\bar{u})\mathscr{O}_{2}^{(j^{3}_{2})}(0)=\sum_{k}C_{12k}\,|u|^{\Delta_{k}-\Delta_{1}-\Delta_{2}}\mathscr{O}_{k}^{(j^{3}_{k})}(0)+\text{descendants}, (65)

with structure constants C12kC_{12k}, and jk3=j13+j23j^{3}_{k}=j^{3}_{1}+j^{3}_{2}.

5.1 The OPE of two interaction operators

The OPE of two interaction operators appears in the limit u1u\to 1 of G(u,u¯)G(u,\bar{u}). To extract this limit from G(x)G(x), we have to find the inverse maps x𝔞(u)x_{\mathfrak{a}}(u) which contribute to the singularities near u=1u=1. For both GR(x)G_{R}(x) and Gσ(x)G_{\sigma}(x), there are clearly only two contributions, i.e. limits where G(x)G(x) becomes singular, namely:888These correspond to the solutions of t1(x)=xt_{1}(x)=x. Fortunately, we do not need to find the other solutions of the 2n2nth-order polynomial equation (43) for u=1u_{*}=1. x=x=\infty and x=1n2x=\frac{1-n}{2}, the former with multiplicity one, and the latter with multiplicity three. We label the two corresponding functions, given in (177), as x𝔞1(u)x^{1}_{\mathfrak{a}}(u), with a (gothic) index 𝔞=1,2\mathfrak{a}=\mathfrak{1,2}, and the superscript indicating that u1u\to 1. Each function gives a channel of the fusion rule, according to Eq.(49). Both functions GR(x)G_{R}(x) and Gσ(x)G_{\sigma}(x) have the same behavior in these limits, as it was necessary for consistency, since both functions should give the same OPE [O[2](int)]×[O[2](int)]=[𝟙]+[σ[3]],[O^{(\mathrm{int})}_{[2]}]\times[O^{(\mathrm{int})}_{[2]}]=[\mathds{1}]+[\sigma_{[3]}], where the r.h.s. is based on the composition of permutations. We mostly focus on GR(x)G_{R}(x) in what follows, similar calculations for Gσ(x)G_{\sigma}(x) are listed in Appendix C.

Determining the constants of integration

For xx\to\infty, GR(x)CRx2G_{R}(x)\approx C_{R}x^{2}. Inserting x=x11(u)x=x_{\mathfrak{1}}^{1}(u) given by Eq.(177), we obtain

GR(x11(u))=CR16n2(1u)2+0×11u+non-singularG_{R}(x_{\mathfrak{1}}^{1}(u))=C_{R}\,{16n^{2}\over(1-u)^{2}}+0\times{1\over 1-u}+\text{non-singular} (66)

By formula (65), since O[2](int)O^{(\mathrm{int})}_{[2]} has weight hint=1h_{int}=1, the leading singular term shows an operator of dimension h=22=0h=2-2=0 — the identity operator. Also, the coefficient of the term (1u)1\sim(1-u)^{-1} is zero, hence there is no contribution from a field of dimension h=1h=1, as it was to be expected for a truly marginal deformation.


The function in Eq.(66) corresponds to a correlator where the permutations in the twists form one representative element of the equivalence class where the 2-cycles of the interaction operators cancel. This happens when they share both elements. At order N1N^{-1}, there must be 𝐬=n+1{\bf s}=n+1 elements entering the permutation, c.f. Eq.(46), so we can take this representative function to be

R(1,,n)()O(1,n+1)(int)(1)O(1,n+1)(int)(u,u¯)R(n,,1)+(0)=GR(x11(u))GR(x11(u¯))\big{\langle}R^{-}_{(1,\cdots,n)}(\infty)O^{(\mathrm{int})}_{(1,n+1)}(1)O^{(\mathrm{int})}_{(1,n+1)}(u,\bar{u})R^{+}_{(n,\cdots,1)}(0)\big{\rangle}=G_{R}(x^{1}_{\mathfrak{1}}(u))G_{R}(x^{1}_{\mathfrak{1}}(\bar{u})) (67)

or any other with a global relabeling of elements in the cycles.

We now fix the constant ainta_{\mathrm{int}} in (53) so that the non-SNS_{N}-invariant two-point functions are normalized,

O2(int)()O2(int)(1)=1.\big{\langle}O^{(\mathrm{int})}_{2}(\infty)O^{(\mathrm{int})}_{2}(1)\big{\rangle}=1. (68)

Note that in these functions the two-cycles must share both of their elements, since, as in Eq.(44), we must have (2)(2)1=1(2)_{\infty}(2)_{1}=1. With this definition, the normalized SNS_{N}-invariant operator is

O[2](int)(z,z¯)=1𝒮2(N)hSNOh1(12)h(int)(z,z¯).O^{(\mathrm{int})}_{[2]}(z,\bar{z})=\frac{1}{\mathscr{S}_{2}(N)}\sum_{h\in S_{N}}O^{(\mathrm{int})}_{h^{-1}(12)h}(z,\bar{z}). (69)

Together with the normalization (15), inserting the limit (66), back into the four-point function (67) we find 16n2CR=116n^{2}C_{R}=1. The same reasoning can be applied to the function Gσ(x)G_{\sigma}(x), which has the exact same limit as (66) in this channel. Therefore

CR=116n2=Cσ.C_{R}=\frac{1}{16n^{2}}=C_{\sigma}. (70)

With the functions G(x)G(x) completely fixed, we can now look at other OPEs and derive structure constants.

The σ3\sigma_{3} channel

In the other channel corresponding to u1u\to 1, we must expand GR(x)G_{R}(x) around x=1n2x=\frac{1-n}{2}, and insert x21(u)x^{1}_{\mathfrak{2}}(u) given by Eq.(177),

GR(x21(u))=4(n21)13(n+1n1)n2(13n)43CR(1u)4/3+15(7+2n2)(n29(n21))13(n+1n1)n2CR(1u)2/34373(n+1n1)n2(n21)13n4CR(1u)1/3+non-singular\begin{split}G_{R}(x^{1}_{\mathfrak{2}}(u))=&-\frac{4(n^{2}-1)^{\frac{1}{3}}\left(\frac{n+1}{n-1}\right)^{\frac{n}{2}}\left(\tfrac{1}{3}n\right)^{\frac{4}{3}}C_{R}}{(1-u)^{4/3}}\\ &+\frac{\tfrac{1}{5}(7+2n^{2})\left(\frac{n^{2}}{9(n^{2}-1)}\right)^{\frac{1}{3}}\left(\frac{n+1}{n-1}\right)^{\frac{n}{2}}C_{R}}{(1-u)^{2/3}}\\ &-\frac{4\cdot 3^{-\frac{7}{3}}\left(\frac{n+1}{n-1}\right)^{\frac{n}{2}}(n^{2}-1)^{\frac{1}{3}}n^{4}C_{R}}{(1-u)^{1/3}}+\text{non-singular}\end{split} (71)

Once again, the coefficient of next-to-leading divergence, (1u)3/3\sim(1-u)^{3/3}, vanishes, showing that there is no dimension-one operator in this conformal family either. The leading singularity shows the presence of an operator of dimension 23=h3σ\tfrac{2}{3}=h^{\sigma}_{3}, so we have found σ3\sigma_{3} itself, and the OPE

O2(int)(u,u¯)O2(int)(1)=O2(int)()σ3(1)O2(int)(0)|1u|8/3σ3(1)+O^{(\mathrm{int})}_{2}(u,\bar{u})O^{(\mathrm{int})}_{2}(1)=\frac{\big{\langle}O^{(\mathrm{int})}_{2}(\infty)\sigma_{3}(1)O^{(\mathrm{int})}_{2}(0)\big{\rangle}}{|1-u|^{8/3}}\,\sigma_{3}(1)+\cdots (72)

whose structure constant is given in Eq.(182), and found independently from Gσ(x)G_{\sigma}(x). Inserting the OPE into the correlation function we find the structure constant

Cn3nRσR+Rn()σ3(1)Rn+(0)C^{R^{-}\sigma R^{+}}_{n3n}\equiv\big{\langle}R^{-}_{n}(\infty)\sigma_{3}(1)R^{+}_{n}(0)\big{\rangle} (73)

involving non-SNS_{N}-invariant Ramond fields and one three-twist. The leading term in Eq.(71) gives us

logCn3nRσR+=(n+23)log(n+1)(n23)log(n1)43logn+43log3+13log2\log C^{R^{-}\sigma R^{+}}_{n3n}=\left(n+\tfrac{2}{3}\right)\log(n+1)-\left(n-\tfrac{2}{3}\right)\log(n-1)-\tfrac{4}{3}\log n+\tfrac{4}{3}\log 3+\tfrac{1}{3}\log 2 (74)

after taking Eq.(182) into account.


The correlation function that gives Eq.(71) lies in an equivalence class where the 2-cycles of the interaction operator share only one element, thus forming σ3\sigma_{3}. The multiplicity 3 of the solution x=1n2x=\frac{1-n}{2} for Eq.(43) implies there are three different equivalence classes with this property. Representative functions for each of those classes are999An elegant and useful way of describing the different classes of permutations with the correct cycle structure and which satisfy Eq.(44) is given in Refs.Pakman:2009zz ; Pakman:2009mi in terms of inequivalent diagrams. The permutations in Eqs.(75a)-(75c) correspond, respectively, to the following diagrams in Ref.Pakman:2009mi : a) the top diagram of Fig.4; b) the second diagram in Fig.4; c) the top diagram of Fig.5.

GR(x21(u))GR(x21(u¯))\displaystyle G_{R}(x^{1}_{\mathfrak{2}}(u))G_{R}(x^{1}_{\mathfrak{2}}(\bar{u}))
=R(1,2,,n1,n)()O(1,n+1)(int)(u,u¯)O(1,2)(int)(1)R(n+1,n,n1,,2)+(0)\displaystyle\qquad=\big{\langle}R^{-}_{(1,2,\cdots,n-1,n)}(\infty)O^{(\mathrm{int})}_{(1,n+1)}(u,\bar{u})O^{(\mathrm{int})}_{(1,2)}(1)R^{+}_{(n+1,n,n-1,\cdots,2)}(0)\big{\rangle} (75a)
=R(1,2,,n1,n)()O(2,n+1)(int)(u,u¯)O(1,n+1)(int)(1)R(n+1,n,n1,,2)+(0)\displaystyle\qquad=\big{\langle}R^{-}_{(1,2,\cdots,n-1,n)}(\infty)O^{(\mathrm{int})}_{(2,n+1)}(u,\bar{u})O^{(\mathrm{int})}_{(1,n+1)}(1)R^{+}_{(n+1,n,n-1,\cdots,2)}(0)\big{\rangle} (75b)
=R(1,2,,n1,n)()O(1,2)(int)(u,u¯)O(2,n+1)(int)(1)R(n+1,n,n1,,2)+(0)\displaystyle\qquad=\big{\langle}R^{-}_{(1,2,\cdots,n-1,n)}(\infty)O^{(\mathrm{int})}_{(1,2)}(u,\bar{u})O^{(\mathrm{int})}_{(2,n+1)}(1)R^{+}_{(n+1,n,n-1,\cdots,2)}(0)\big{\rangle} (75c)

One can check that the permutations do satisfy Eq.(44). Note that, by necessity, the twists in the Ramond fields are not the inverse of one another, so the two-point function

R(1,2,,n1,n)()R(n+1,n,n1,,2)+(0)=0.\big{\langle}R^{-}_{(1,2,\cdots,n-1,n)}(\infty)R^{+}_{(n+1,n,n-1,\cdots,2)}(0)\big{\rangle}=0. (76)

Thus we see that the σ3\sigma_{3} channel of the fusion [O[2](int)]×[O[2](int)][O^{(\mathrm{int})}_{[2]}]\times[O^{(\mathrm{int})}_{[2]}] is always present, because the interaction operator is necessarily an SNS_{N}-invariant object, but Eqs.(75) and (76) mean that σ3\sigma_{3} does not contribute to the correction of the two-point functions of individual, non-SNS_{N}-invariant Ramond fields Rn±R^{\pm}_{n}. It only contributes to the SNS_{N}-invariant combination R[n]±R^{\pm}_{[n]}, by weaving together different individual terms.

The OPE RnRn+R^{-}_{n}R^{+}_{n}

Although the positions of the Ramond fields are fixed in (49), we can loosen the punctures back to Eq.(21), fix them differently as z2=z_{2}=\infty, z3=0z_{3}=0, z4=1z_{4}=1, in which case z1=uz_{1}=u, to find

O[2](int)()R[n](u,u¯)O[2](int)(0)R[n]+(1)=|1u|4nGR(u,u¯).\big{\langle}O^{(\mathrm{int})}_{[2]}(\infty)R^{-}_{[n]}(u,\bar{u})O^{(\mathrm{int})}_{[2]}(0)R^{+}_{[n]}(1)\big{\rangle}=|1-u|^{4-n}G_{R}(u,\bar{u}). (77)

Now the limit u1u\to 1 corresponds to the OPE R[n](u,u¯)R[n]+(1)R^{-}_{[n]}(u,\bar{u})R^{+}_{[n]}(1). The expansion near u=1u=1 for channel (66) is

O2(int)()Rn(u,u¯)O2(int)(0)Rn+(1)=1|1u|n+\big{\langle}O^{(\mathrm{int})}_{2}(\infty)R^{-}_{n}(u,\bar{u})O^{(\mathrm{int})}_{2}(0)R^{+}_{n}(1)\big{\rangle}=\frac{1}{|1-u|^{n}}+\cdots (78)

This corresponds to an operator of dimension zero, and is in fact the correct expression for the two-point function of Ramond fields, Eq.(15). In the channel (71) we now find the behavior (1u)n+83\sim(1-u)^{-n+\frac{8}{3}}, indicating a twist-three operator of holomorphic weight

h=n+42+h3σ.h=\frac{n+4}{2}+h^{\sigma}_{3}. (79)

To understand the appearance of σ3\sigma_{3} in a channel of the OPE RnRn+R^{-}_{n}R^{+}_{n}, let us consider the simpler case of the correlator with bare twists only. Changing the points of Eq.(33), we can find the OPE σnσn\sigma_{n}\sigma_{n} from the limit u1u\to 1 of the function

σ2()σn(1)σn(u,u¯)σ2(0)=|1u|4(h2σhnσ)g(u,u¯).\big{\langle}\sigma_{2}(\infty)\sigma_{n}(1)\sigma_{n}(u,\bar{u})\sigma_{2}(0)\big{\rangle}=|1-u|^{4(h^{\sigma}_{2}-h^{\sigma}_{n})}g(u,\bar{u}). (80)

Channel (183) gives an operator of dimension zero, and channel (184) an operator of dimension 23=h3σ\frac{2}{3}=h^{\sigma}_{3}. This gives us the fusion rule

[σn]×[σn]=[𝟙]+[σ3]+[\sigma_{n}]\times[\sigma_{n}]=[{\mathds{1}}]+[\sigma_{3}]+\cdots (81)

Of course, there are other twists in the r.h.s. but they cannot be found from the four-point function we have began with, because of the condition (44). As discussed above, in channel (183) the two twists σ2\sigma_{2} in the correlator have inverse cycles, hence it is necessary that the two twists σn\sigma_{n} also be the inverse of each other; this gives 𝟙\mathds{1} in the fusion rule. As for the channel (184), we have seen that the cycles in σ2\sigma_{2} then only have one overlapping element, say, σ(k)σ(km)=σ(km)\sigma_{(k\ell)}\sigma_{(km)}=\sigma_{(k\ell m)}. Hence for Eq.(44) to be satisfied the two σn\sigma_{n} operators must compose to σnσn=σ(mk)\sigma_{n}\sigma_{n}=\sigma_{(m\ell k)}, which is why σ3\sigma_{3} appears.

5.2 Non-BPS operators in the OPEs of Rn±R^{\pm}_{n} with O2(int)O^{(\mathrm{int})}_{2}

We now turn to the limit u0u\to 0, where the interaction operator collides with either the Ramond field Rn+(0)R^{+}_{n}(0) or with the bare twisted field σn(0)\sigma_{n}(0), depending on the function we analyze, if either GRG_{R} or GσG_{\sigma}. Now one can find all 2n2n solutions of Eq.(43), viz. x=0x=0 (with multiplicity n1n-1) and x=nx=-n (with multiplicity n+1n+1), all contributing to the OPE limits.


The function GR(u,u¯)G_{R}(u,\bar{u}) gives the OPE O[2](int)(u,u¯)R[n]+(0)O^{(\mathrm{int})}_{[2]}(u,\bar{u})R^{+}_{[n]}(0). Using (165),

GR(x10(u))\displaystyle G_{R}(x^{0}_{\mathfrak{1}}(u)) =u54n2n1(C1+c1u1n1+),\displaystyle=u^{-\frac{5}{4}\frac{n-2}{n-1}}\big{(}C_{\mathfrak{1}}+c_{\mathfrak{1}}u^{1\over n-1}+\cdots\big{)}, (82)
GR(x20(u))\displaystyle G_{R}(x^{0}_{\mathfrak{2}}(u)) =u143n2n+1(C2+c2u1n+1+).\displaystyle=u^{-\frac{1}{4}\frac{3n-2}{n+1}}\big{(}C_{\mathfrak{2}}+c_{\mathfrak{2}}u^{1\over n+1}+\cdots\big{)}. (83)

Counting powers of uu, we find that the OPE O2(int)Rn+O^{(\mathrm{int})}_{2}R^{+}_{n} results in a twisted field Ym+Y^{+}_{m} which is positively R-charged with j3=12j^{3}=\frac{1}{2}, and has the holomorphic dimension

hmY=32m+hmσ.h^{Y}_{m}=\frac{3}{2m}+h^{\sigma}_{m}. (84)

Channels x10(u)x^{0}_{\mathfrak{1}}(u) and x20(u)x^{0}_{\mathfrak{2}}(u) give m=n1m=n-1 and m=n+1m=n+1, respectively.

The OPE O[2](int)RnO^{(\mathrm{int})}_{[2]}R^{-}_{n}, is obtained in the limit u0u\to 0 of

O[2](int)()R[n](u,u¯)O[2](int)(0)R[n]+(1)=|1u|4nGR(u,u¯),\big{\langle}O^{(\mathrm{int})}_{[2]}(\infty)R^{-}_{[n]}(u,\bar{u})O^{(\mathrm{int})}_{[2]}(0)R^{+}_{[n]}(1)\big{\rangle}=|1-u|^{4-n}G_{R}(u,\bar{u}), (85)

which follows from the same procedure of fixing points used to find (77). Since the factor of (1u)4n2(1-u)^{\frac{4-n}{2}} does not contribute to the leading term near u=0u=0, we immediately find the same expansion as before. Now the resulting fields YmY^{-}_{m} have the same dimensions (84), but opposite R-charge, j3=12j^{3}=-\frac{1}{2}.

In summary, we have found the fusion rules

[O[2](int)]×[R[n]±]=[Y[n1]±]+[Y[n+1]±],[O^{(\mathrm{int})}_{[2]}]\times[R^{\pm}_{[n]}]=[Y^{\pm}_{[n-1]}]+[Y^{\pm}_{[n+1]}], (86)

where the fields Ym±Y^{\pm}_{m} have the dimension (84). The appearance of m=n±1m=n\pm 1 in the r.h.s. is a basic consequence of permutation composition, see Eq.(186). We take the Ym±Y^{\pm}_{m} to be normalized, so that (by charge conservation) the non-vanishing two-point functions are

Ym±(z1,z¯1)Ym(z2,z¯2)=|z12|4hmY.\big{\langle}Y^{\pm}_{m}(z_{1},\bar{z}_{1})Y^{\mp}_{m}(z_{2},\bar{z}_{2})\big{\rangle}=|z_{12}|^{-4h^{Y}_{m}}.

Inserting the OPE back into the four-point function, the leading short-distance coefficients C𝔞C_{\mathfrak{a}} give us information about the product of structure constants

Rn()O2(int)(1)Yn1+(0)Rn+()O2(int)(1)Yn1(0)\displaystyle\big{\langle}R^{-}_{n}(\infty)O^{(\mathrm{int})}_{2}(1)Y^{+}_{n-1}(0)\big{\rangle}\,\big{\langle}R^{+}_{n}(\infty)O^{(\mathrm{int})}_{2}(1)Y^{-}_{n-1}(0)\big{\rangle} =|C1|2,\displaystyle=|C_{\mathfrak{1}}|^{2}, (87a)
Rn()O2(int)(1)Yn+1+(0)Rn+()O2(int)(1)Yn+1(0)\displaystyle\big{\langle}R^{-}_{n}(\infty)O^{(\mathrm{int})}_{2}(1)Y^{+}_{n+1}(0)\big{\rangle}\,\big{\langle}R^{+}_{n}(\infty)O^{(\mathrm{int})}_{2}(1)Y^{-}_{n+1}(0)\big{\rangle} =|C2|2\displaystyle=|C_{\mathfrak{2}}|^{2} (87b)

(Recall that we must take |GR(x𝔞0(u))|2|G_{R}(x^{0}_{\mathfrak{a}}(u))|^{2}.) In the l.h.s. we actually have products of conjugate three-point functions/structure constants,

Cn,2,mRO(int)Y±=RnO2(int)Ym±=Cn,2,mR±O(int)Y,C^{R^{\mp}O^{(\mathrm{int})}Y^{\pm}}_{n,2,m}=\big{\langle}R^{\mp}_{n}O^{(\mathrm{int})}_{2}Y^{\pm}_{m}\big{\rangle}=C^{R^{\pm}O^{(\mathrm{int})}Y^{\mp}}_{n,2,m}, (88)

(with the twists in the subscripts) and taking the explicit expressions for C𝔞C_{\mathfrak{a}}, found from the expansions (82)-(83), we get

logCn,2,n1R±O(int)Y\displaystyle\log C^{R^{\pm}O^{(\mathrm{int})}Y^{\mp}}_{n,2,n-1} =n+22log(n1)+n24n22(n1)logn,\displaystyle=-\frac{n+2}{2}\log(n-1)+\frac{n^{2}-4n-2}{2(n-1)}\log n, (89)
logCn,2,n+1R±O(int)Y\displaystyle\log C^{R^{\pm}O^{(\mathrm{int})}Y^{\mp}}_{n,2,n+1} =+n22log(n+1)n2+4n22(n+1)logn.\displaystyle=+\frac{n-2}{2}\log(n+1)-\frac{n^{2}+4n-2}{2(n+1)}\log n. (90)

A similar situation takes place in the OPE O2(int)σnO^{(\mathrm{int})}_{2}\sigma_{n}, found in the limit u0u\to 0 of Gσ(u,u¯)G_{\sigma}(u,\bar{u}). The two channels reveal twisted operators 𝒴m\mathcal{Y}_{m} with zero R-charge and dimensions

hm𝒴=58(m2m)+hmσ,h^{\mathcal{Y}}_{m}=-\frac{5}{8}\left(\frac{m-2}{m}\right)+h^{\sigma}_{m}, (91)

in a fusion rule

[O[2](int)]×[σ[n]]=[𝒴[n1]]+[𝒴[n+1]][O^{(\mathrm{int})}_{[2]}]\times[\sigma_{[n]}]=[\mathcal{Y}_{[n-1]}]+[\mathcal{Y}_{[n+1]}] (92)

the terms in the r.h.s. corresponding to x10(u)x^{0}_{\mathfrak{1}}(u) and x20(u)x^{0}_{\mathfrak{2}}(u), respectively. The coefficients calculated from the expansion of GσG_{\sigma} give structure constants as before:

logCn,2,n1σO(int)𝒴\displaystyle\log C^{\sigma O^{(\mathrm{int})}\mathcal{Y}}_{n,2,n-1} =(n+1)22nlog(n1)+n24n12(n1)logn\displaystyle=-\frac{(n+1)^{2}}{2n}\log(n-1)+\frac{n^{2}-4n-1}{2(n-1)}\log n (93)
logCn,2,n+1σO(int)𝒴\displaystyle\log C^{\sigma O^{(\mathrm{int})}\mathcal{Y}}_{n,2,n+1} =+(n1)22nlog(n+1)n2+4n12(n+1)logn\displaystyle=+\frac{(n-1)^{2}}{2n}\log(n+1)-\frac{n^{2}+4n-1}{2(n+1)}\log n (94)

where σn()O2(int)(1)𝒴m(0)Cn,2,mσO(int)𝒴\big{\langle}\sigma_{n}(\infty)O^{(\mathrm{int})}_{2}(1)\mathcal{Y}_{m}(0)\big{\rangle}\equiv C^{\sigma O^{(\mathrm{int})}\mathcal{Y}}_{n,2,m}.


We have found that the operator algebras of Ramond fields with the deformation operator include non-BPS fields. These fields are consistent with the fractional spectral flow with ξ=nn+1\xi=\frac{n}{n+1} of twisted non-BPS fields in the NS sector, recently found deBeer:2019ioe to be a part of the OPEs of the deformation operator and NS chiral operators; see the discussion in §6.5. A complete study of the algebras found here requires knowledge of OPEs such as [Yn±1±]×[O[2](int)][Y^{\pm}_{n\pm 1}]\times[O^{(\mathrm{int})}_{[2]}]. For that, the new fields have to be explicitly constructed. From our discussion of their properties, and in particular from the conformal dimension (84), we can infer that

Ym±(0)=G12m±J1m3σm(0).Y^{\pm}_{m}(0)=G^{\pm}_{-\frac{1}{2m}}J^{3}_{-\frac{1}{m}}\sigma_{m}(0). (95)

This explicit construction should be sufficient for the study of the remaining OPEs by the computation of four-point functions such as Y[m]()O[2](int)(1)O[2](int)(u,u¯)Y[m]+\langle Y^{-}_{[m]}(\infty)O^{(\mathrm{int})}_{[2]}(1)O^{(\mathrm{int})}_{[2]}(u,\bar{u})Y^{+}_{[m]}\rangle (which, incidentally, can be computed with the same covering map used here).

6 Analytic regularization and field renormalization

We now turn to the calculation of the conformal dimension of Ramond fields in the interacting SCFT2. At second-order in perturbation theory, this requires computation of the integral (24), using the functions we have found in Sect.4.

6.1 Dotsenko-Fateev integrals

We want to compute integrals J=d2uG(u,u¯)J=\int d^{2}u\,G(u,\bar{u}), given an analytic expression for

G(x)=Cxα1(x1)α2(x+n)α3(x+n1)α4(x+n12)α5,with{α1+n2=α4nα2n2=α3+nG(x)=C\,\frac{x^{\alpha_{1}}(x-1)^{\alpha_{2}}(x+n)^{\alpha_{3}}(x+n-1)^{\alpha_{4}}}{(x+\frac{n-1}{2})^{\alpha_{5}}},\quad\text{with}\quad\begin{cases}\alpha_{1}+n-2=\alpha_{4}-n\\ \alpha_{2}-n-2=\alpha_{3}+n\end{cases} (96)

from which G(u,u¯)G(u,\bar{u}) is obtained by inversion of the map (32). Both the Ramond function (61) and the bare twist function (64) have the form (96). We can perform a change of variables from uu to xx in the integral which, taking the special relation between the exponents into account, becomes

J\displaystyle J =d2x|u(x)G(x)|2=(4nC)2d2x|[x(x+n1)]α4n[(x1)(x+n)]α3+n(x+n12)α52|2.\displaystyle=\int\!d^{2}x\,|u^{\prime}(x)G(x)|^{2}=(4nC)^{2}\int\!d^{2}x\left|\frac{\left[x(x+n-1)\right]^{\alpha_{4}-n}\left[(x-1)(x+n)\right]^{\alpha_{3}+n}}{(x+\frac{n-1}{2})^{\alpha_{5}-2}}\right|^{2}. (97)

We then make the following change of variables Pakman:2009mi ,

y(x)=4(x1)(x+n)(n+1)2,y(x)=-\frac{4(x-1)(x+n)}{(n+1)^{2}}, (98)

such that every term in the new integrand is expressed simply in terms of yy,

J(n)=(4nC)24(n+12)4[a+b+c+1]I(n),\displaystyle J(n)=\frac{(4nC)^{2}}{4}\Big{(}\frac{n+1}{2}\Big{)}^{4\left[a+b+c+1\right]}\;I(n), (99)
I(n)d2y|y|2a|1y|2b|ywn|2c,\displaystyle I(n)\equiv\int\!d^{2}y\;|y|^{2a}|1-y|^{2b}|y-w_{n}|^{2c}, (100)

where

a=α3+n,b=α512,c=α4n,wn=4n(n+1)2.a=\alpha_{3}+n,\quad b=-\frac{\alpha_{5}-1}{2},\quad c=\alpha_{4}-n,\quad w_{n}=\frac{4n}{(n+1)^{2}}. (101)

We will refer to I(n)I(n) as a ‘Dotsenko-Fateev (DF) integral’, as it has been studied in detail by Dotsenko and Fateev, as a representation of correlation functions in degenerate CFTs Dotsenko:1984nm ; Dotsenko:1984ad ; dotsenko1988lectures .101010Cf. also Refs.Mussardo:1987eq ; Mussardo:1987ab ; Mussardo:1988av .

The properties of I(n)I(n) crucially depend on the exponents of its critical points y={0,wn,1}y=\{0,w_{n},1\}. For example, the exponents for GR(x)G_{R}(x) are

aR=12+14n,bR=32,cR=1214na_{R}=\tfrac{1}{2}+\tfrac{1}{4}n,\qquad b_{R}=-\tfrac{3}{2},\qquad c_{R}=\tfrac{1}{2}-\tfrac{1}{4}n (102)

thus, for general nn, all three critical points are branching points. The integral diverges at 11 and \infty, and vanishes at 0 for all nn; at wnw_{n}, it converges for n6n\leq 6 and diverges for n>6n>6. Clearly, some regularization procedure is needed. Following Ref.dotsenko1988lectures , we now show that I(n)I(n) can be expressed in terms of hypergeometric functions, leading to a regularization by analytic continuation. We do this in two steps:

  1. 1.

    Assume that the parameters a,b,ca,b,c are such that the DF integral exists.

  2. 2.

    Express the integrals in terms of an analytic function of a,b,ca,b,c that is well-defined also for values of a,b,ca,b,c, such as (102), for which the original integral diverges. (Such functions will turn out to be a product of hypergeometric and Gamma functions.) This leads to an extension of the definition of the integrals by their maximal analytic continuation.

As we shall see, the procedure is consistent.

Let us write yy\in\mathbb{C} as y=y1+iy2y=y_{1}+iy_{2} in (100), and perform a rotation of y2y_{2}, such that y2i(12iε)y2y_{2}\mapsto i(1-2i\varepsilon)y_{2}, with ε\varepsilon a positive arbitrarily small parameter. Defining v±=y1±y2v_{\pm}=y_{1}\pm y_{2} (where y2y_{2} now refers to the new, rotated coordinate), and expanding the integrand to first order in ε\varepsilon,

I\displaystyle I =i𝑑y1𝑑y2[y12y22(14iε)]a[(y11)2y22(14iε)]b\displaystyle=i\iint dy_{1}dy_{2}\;\left[y_{1}^{2}-y_{2}^{2}(1-4i\varepsilon)\right]^{a}\big{[}(y_{1}-1)^{2}-y_{2}^{2}(1-4i\varepsilon)\big{]}^{b}
×[(y1wn)2y22(14iε)]c\displaystyle\hskip 76.82234pt\times\big{[}(y_{1}-w_{n})^{2}-y_{2}^{2}(1-4i\varepsilon)\big{]}^{c}
=i2𝑑v𝑑v+[{viε(vv+)}{v++iε(vv+)}]a\displaystyle=\frac{i}{2}\iint dv_{-}dv_{+}\;\left[\left\{v_{-}-i\varepsilon(v_{-}-v_{+})\right\}\left\{v_{+}+i\varepsilon(v_{-}-v_{+})\right\}\right]^{a}
×[{v1iε(vv+)}{v+1+iε(vv+)}]b\displaystyle\hskip 76.82234pt\times\left[\left\{v_{-}-1-i\varepsilon(v_{-}-v_{+})\right\}\left\{v_{+}-1+i\varepsilon(v_{-}-v_{+})\right\}\right]^{b}
×[{vwniε(vv+)}{v+wn+iε(vv+)}]c.\displaystyle\hskip 76.82234pt\times\left[\left\{v_{-}-w_{n}-i\varepsilon(v_{-}-v_{+})\right\}\left\{v_{+}-w_{n}+i\varepsilon(v_{-}-v_{+})\right\}\right]^{c}.

The double integrals have been factorized into a product of two one-dimensional integrals,

I=i2+𝑑v[viε(vv+)]a[v1iε(vv+)]b×[vwniε(vv+)]c×+dv+[v++iε(vv+)]a[v+1+iε(vv+)]b×[v+wn+iε(vv+)]c\begin{split}I=\frac{i}{2}\int_{-\infty}^{+\infty}\!dv_{-}&\left[v_{-}-i\varepsilon(v_{-}-v_{+})\right]^{a}\left[v_{-}-1-i\varepsilon(v_{-}-v_{+})\right]^{b}\\ \times&\left[v_{-}-w_{n}-i\varepsilon(v_{-}-v_{+})\right]^{c}\\ \quad\times\int_{-\infty}^{+\infty}\!dv_{+}\;&\left[v_{+}+i\varepsilon(v_{-}-v_{+})\right]^{a}\left[v_{+}-1+i\varepsilon(v_{-}-v_{+})\right]^{b}\\ \times&\left[v_{+}-w_{n}+i\varepsilon(v_{-}-v_{+})\right]^{c}\end{split} (103)

because the variable v±v_{\pm} only appears in the vv_{\mp} integral multiplied by the infinitesimal parameter ε\varepsilon. The effect of the ε\varepsilon-terms is to specify how the otherwise real integrals of

f(ζ)=ζa(ζ1)b(ζwn)c,ζ,wn(0,1),f(\zeta)=\zeta^{a}(\zeta-1)^{b}(\zeta-w_{n})^{c}\ ,\qquad\zeta\in{\mathbb{C}}\ ,\qquad w_{n}\in(0,1)\subset{\mathbb{R}}, (104)

go around the points 0,wn,10,w_{n},1. To further disentangle the integrals, we split integration over v+v_{+} at 0,wn,10,w_{n},1, so that ε\varepsilon-terms can be ignored, while the vv_{-} integrals go around the contours γk\gamma_{k} dictated by the infinitesimal terms ε(vv+)\varepsilon(v_{-}-v_{+}) as in Fig.1(a),

I=i2[0𝑑v+f(v+)γ0𝑑vf(v)+0wn𝑑v+f(v+)γ1𝑑vf(v)+wn1dv+f(v+)γ2dvf(v)+1dv+f(v+)γ3dvf(v)].\begin{split}I=\frac{i}{2}\Bigg{[}&\int_{-\infty}^{0}dv_{+}f(v_{+})\int_{\gamma_{0}}dv_{-}f(v_{-})+\int_{0}^{w_{n}}dv_{+}f(v_{+})\int_{\gamma_{1}}dv_{-}f(v_{-})\\ +&\int_{w_{n}}^{1}dv_{+}f(v_{+})\int_{\gamma_{2}}dv_{-}f(v_{-})+\int_{1}^{\infty}dv_{+}f(v_{+})\int_{\gamma_{3}}dv_{-}f(v_{-})\Bigg{]}.\end{split} (105)

For example, for v+(0,wn)v_{+}\in(0,w_{n}),

ε(vv+)|v=0>0,ε(vv+)|v=wn<0,ε(vv+)|v=1<0,\varepsilon(v_{-}-v_{+})|_{v_{-}=0}>0,\qquad\varepsilon(v_{-}-v_{+})|_{v_{-}=w_{n}}<0,\qquad\varepsilon(v_{-}-v_{+})|_{v_{-}=1}<0,

hence the contour γ1\gamma_{1} goes above v=0v_{-}=0, and below v=wn,1v_{-}=w_{n},1.

Refer to caption
Figure 1: (a) Contours for the Dotsenko-Fateev integral; (b) Closing γ1\gamma_{1} and γ2\gamma_{2}; (c) Deformation; (d) Final contours (ignoring circles at infinity). Red lines indicate branching cuts.

The function f(ζ)f(\zeta) has branch cuts, so closing the contours γk\gamma_{k} with semi-circles is non-trivial. Here is the point were our regularization procedure effectively starts. Assume that a,b,ca,b,c are such that the DF integral is convergent. Precisely, assume that

a>1,b>1,c>1,1+a+b+c<0,a>-1,\quad b>-1,\quad c>-1,\quad 1+a+b+c<0, (106)

which ensures, respectively, convergence at the points 0,1,wn,0,1,w_{n},\infty. Now close the contours γk\gamma_{k} by making semicircles of radius RR\to\infty on the lower or upper plane. The curve γ0\gamma_{0} passes below every branch point, hence close the contour with a clockwise semicircle Γ\Gamma^{-} on the lower plane; there are no poles inside γ0Γ\gamma_{0}\cup\Gamma^{-}, and limRΓ𝑑vf(v)0\lim_{R\to\infty}\int_{\Gamma^{-}}dv_{-}\ f(v_{-})\to 0 given our assumptions (106); hence γ0𝑑vf(v)=0\int_{\gamma_{0}}dv_{-}f(v_{-})=0. Similarly, γ3𝑑vf(v)=0\int_{\gamma_{3}}dv_{-}f(v_{-})=0, now with the contour on the upper plane. Thus only two terms remain in Eq.(105).

If we try to close γ1\gamma_{1} or γ2\gamma_{2} in Fig.1(a), we are deemed to cross branch cuts, and move to another Riemann sheet of f(v)f(v_{-}). One way out of this is to cross the cut on a branching point, where f(v)f(v_{-}) is single-valued. That the integral exists at the branching points is assured by our assumptions (106). Thus we choose the branch cuts to align with the Real axis in two different ways: for the integral over γ1\gamma_{1} they extend to -\infty, and for γ2\gamma_{2} they extend to ++\infty; then we close the contours with semi-circles as in Fig.1(b). In one case, we cross the real axis at v=0v_{-}=0, in the other at v=wnv_{-}=w_{n}. Next, we deform the contours as in Fig.1(c). Given our assumptions (106), as RR\to\infty the integral over the (almost closed) circle vanishes, and we have γi𝑑vf(v)=Ci𝑑vf(v)\int_{\gamma_{i}}dv_{-}f(v_{-})=\int_{C_{i}}dv_{-}f(v_{-}) for i=1,2i=1,2, where the contours CiC_{i} are shown in Fig.1(d). Integration over CiC_{i} is standard: the effect of coasting the two margins of a branch cut, turning at the branch point is to produce a phase 2isin(πθ)2i\sin(\pi\theta).

Thus we arrive at the following form of (105),

I(a,b,c;wn)=s(a)I~1(a,b,c;wn)I2(a,b,c;wn)s(b)I1(a,b,c;wn)I~2(a,b,c;wn),\begin{split}I(a,b,c;w_{n})=&-s(a)\tilde{I}_{1}(a,b,c;w_{n})\ I_{2}(a,b,c;w_{n})\\ &-s(b)I_{1}(a,b,c;w_{n})\ \tilde{I}_{2}(a,b,c;w_{n}),\end{split} (107)

where s(θ)sin(πθ)s(\theta)\equiv\sin(\pi\theta) and we have defined four ‘canonical integrals’:

I1(a,b,c;wn)\displaystyle I_{1}(a,b,c;w_{n}) 1𝑑v+v+a(v+1)b(v+wn)c\displaystyle\equiv\int_{1}^{\infty}dv_{+}\,v_{+}^{a}(v_{+}-1)^{b}(v_{+}-w_{n})^{c} (108a)
I2(a,b,c;wn)\displaystyle I_{2}(a,b,c;w_{n}) 0wn𝑑vva(1v)b(wnv)c\displaystyle\equiv\int_{0}^{w_{n}}dv_{-}\,v_{-}^{a}(1-v_{-})^{b}(w_{n}-v_{-})^{c} (108b)
I~1(a,b,c;wn)\displaystyle\tilde{I}_{1}(a,b,c;w_{n}) 0𝑑v+(v+)a(1v+)b(wnv+)c\displaystyle\equiv\int_{-\infty}^{0}dv_{+}\,(-v_{+})^{a}(1-v_{+})^{b}(w_{n}-v_{+})^{c} (108c)
I~2(a,b,c;wn)\displaystyle\tilde{I}_{2}(a,b,c;w_{n}) wn1𝑑vva(1v)b(vwn)c\displaystyle\equiv\int_{w_{n}}^{1}dv_{-}\,v_{-}^{a}(1-v_{-})^{b}(v_{-}-w_{n})^{c} (108d)

The I~1,2\tilde{I}_{1,2} can actually be written in terms of the I1,2I_{1,2} with a different arrangement of their arguments:

I~1(a,b,c;wn)=I1(b,a;c;1wn),I~2(a,b,c;wn)=I2(b,a;c;1wn).\tilde{I}_{1}(a,b,c;w_{n})=I_{1}(b,a;c;1-w_{n}),\quad\tilde{I}_{2}(a,b,c;w_{n})=I_{2}(b,a;c;1-w_{n}). (109)

Also, by combining deformed contours such as the ones in Fig.1, it can be shown Dotsenko:1984nm that I1,2(a,b,c,wn)I_{1,2}(a,b,c,w_{n}) and I~1,2(a,b,c,wn)\tilde{I}_{1,2}(a,b,c,w_{n}), with the same arguments, form a linear system:

s(b+c)I1=s(a)I~1s(c)I~2,s(b+c)I2=s(a+b+c)I~1s(b)I~2.s(b+c)I_{1}=s(a)\tilde{I}_{1}-s(c)\tilde{I}_{2}\,,\quad s(b+c)I_{2}=-s(a+b+c)\tilde{I}_{1}-s(b)\tilde{I}_{2}. (110)

The four canonical integrals are proportional to the Euler representation of the hypergeometric function NIST:DLMF151 ,

01𝑑ttβ1(1wt)α(1t)γβ1=Γ(β)Γ(γβ)Γ(γ)F(α,β;γ;w)valid for|arg(1w)|<π,0<Re(β)<Re(γ).\displaystyle\begin{split}&\int_{0}^{1}\!dt\,t^{\beta-1}(1-wt)^{-\alpha}(1-t)^{\gamma-\beta-1}=\frac{\Gamma(\beta)\Gamma(\gamma-\beta)}{\Gamma(\gamma)}F(\alpha,\beta;\gamma;w)\\ &\text{valid for}\quad|\mathrm{arg}(1-w)|<\pi,\quad 0<\mathrm{Re}(\beta)<\mathrm{Re}(\gamma).\end{split} (111)

With the substitution t=1/v+t=1/v_{+} in I1I_{1}, and t=v/wnt=v_{-}/w_{n} in I2I_{2}, we find

I1(a,b;c;wn)\displaystyle I_{1}(a,b;c;w_{n}) =Γ(1abc)Γ(1+b)Γ(ac)F(c,1abc;ac;wn),\displaystyle=\frac{\Gamma(-1-a-b-c)\Gamma(1+b)}{\Gamma(-a-c)}F(-c,-1-a-b-c;-a-c;w_{n}), (112)
I2(a,b,c;wn)\displaystyle I_{2}(a,b,c;w_{n}) =Γ(1+a)Γ(1+c)Γ(2+a+c)wn1+a+cF(b,1+a;2+a+c;wn).\displaystyle=\frac{\Gamma(1+a)\Gamma(1+c)}{\Gamma(2+a+c)}w_{n}^{1+a+c}F(-b,1+a;2+a+c;w_{n}). (113)

The restrictions (111), required for both integrals to be represented by hypergeometrics, translate to a,b,ca,b,c as

1<a<1,1<b<a,1<c<1ab,-1<a<1,\quad-1<b<-a,\quad-1<c<-1-a-b, (114)

and also 0<wn<10<w_{n}<1, cf. (101). These conditions are consistent with our starting hypothesis (106), therefore Eq.(107) can be read as a product of hypergeometric and Gamma functions.

The ‘canonical functions’ (112) and (113) are analytic functions of each of the parameters a,b,ca,b,c, on the domain of validity (114). This is evident for the Gamma functions, and is also true for the hypergeometrics, see (bateman1953higher, , §2.1.6). Note that in (112) and (113) what actually appears is the ‘regularized hypergeometric function’

𝐅(α,β;γ;w)1Γ(γ)F(α,β;γ;w),{\bf F}(\alpha,\beta;\gamma;w)\equiv\frac{1}{\Gamma(\gamma)}F(\alpha,\beta;\gamma;w), (115)

which is an entire function of α,β,γ\alpha,\beta,\gamma (bateman1953higher, , §2.1.6). In particular, 𝐅(α,β;γ;w){\bf F}(\alpha,\beta;\gamma;w) is regular at γ=k\gamma=-k, with kk\in\mathbb{N}, where the Gamma function develops a pole and (NIST:DLMF, , §15.2)

𝐅(α,β;k;w)=Γ(α+k+1)Γ(β+k+1)Γ(α)Γ(β)(k+1)!wk+1F(α+k+1,β+k+1;k+2;w).{\bf F}(\alpha,\beta;-k;w)=\frac{\Gamma(\alpha+k+1)\Gamma(\beta+k+1)}{\Gamma(\alpha)\Gamma(\beta)(k+1)!}w^{k+1}F(\alpha+k+1,\beta+k+1;k+2;w). (116)

Hence I(a,b,c;wn)I(a,b,c;w_{n}) is analytic in a,b,ca,b,c separately. Consequently, an analytic continuation of I(a,b,c;wn)I(a,b,c;w_{n}) to outside of the domain of definition (114) is unique, when it exists. We take this analytic continuation to be the definition of the DF integral (100) for arbitrary parameters. Note that it is not precluded that, outside the domain (114), I(a,b,c;wn)I(a,b,c;w_{n}) might develop a singularity — there may be a barrier to the analytic continuation — it just happens that, for the applications below, the continuation is, indeed, (almost) always well-defined.

6.2 The integral for R-charged Ramond fields

Let us apply our results to the Ramond function (61). As noted before, the parameters (102) do not lie within the domain (114), hence we are indeed using the analytic continuation. Eqs.(112), (113), (109) yield

I1(aR,bR,cR)\displaystyle I_{1}(a_{R},b_{R},c_{R}) =π(4n2)32w2F(32,32+14n;3;wn)\displaystyle=\frac{\pi(4-n^{2})}{32}\,w^{2}\,F(\tfrac{3}{2},\tfrac{3}{2}+\tfrac{1}{4}n;3;w_{n}) (117a)
I2(aR,bR,cR)\displaystyle I_{2}(a_{R},b_{R},c_{R}) =1s(12n4)I1(aR,bR,cR)\displaystyle=\frac{1}{s(\tfrac{1}{2}-\tfrac{n}{4})}I_{1}(a_{R},b_{R},c_{R}) (117b)
I~1(aR,bR,cR)\displaystyle\tilde{I}_{1}(a_{R},b_{R},c_{R}) =2πΓ(32+n4)Γ(1+n4)F(12+n4,12;1+n4;1wn)\displaystyle=-\frac{2\sqrt{\pi}\Gamma(\tfrac{3}{2}+\tfrac{n}{4})}{\Gamma(1+\tfrac{n}{4})}F(-\tfrac{1}{2}+\tfrac{n}{4},-\tfrac{1}{2};1+\tfrac{n}{4};1-w_{n}) (117c)
I~2(aR,bR,cR)\displaystyle\tilde{I}_{2}(a_{R},b_{R},c_{R}) =2πΓ(32n4)Γ(1n4)(1w)n/4F(12n4,12;1n4;1wn)\displaystyle=-\frac{2\sqrt{\pi}\Gamma(\tfrac{3}{2}-\tfrac{n}{4})}{\Gamma(1-\tfrac{n}{4})}(1-w)^{-n/4}F(-\tfrac{1}{2}-\tfrac{n}{4},-\tfrac{1}{2};1-\tfrac{n}{4};1-w_{n}) (117d)

Several observations are in order. The expression (117a) does not correspond immediately to the formula (112), because here we have Γ(1)\Gamma(-1) in the denominator. In this case, we must use Eq.(116) to find the correct expression for I1I_{1} in (117a). Expression (117b) can be found immediately from (113). The factor s(c)s(c) in (117b) can be found either from Γ(z)Γ(1z)=π/sin(πz)\Gamma(z)\Gamma(1-z)=\pi/\sin(\pi z), or from the linear system (110), by noting that in the present case we have

s(aR+cR)=0,s(aR)=s(cR),s(aR+bR+cR)=s(bR)=1.s(a_{R}+c_{R})=0,\qquad s(a_{R})=s(c_{R}),\qquad s(a_{R}+b_{R}+c_{R})=-s(b_{R})=-1. (118)

Eqs.(117c) and (119d) follow immediately from (109), but (119d) is only valid when nn is odd. For nn even there are two cases. When n=4(k+1)n=4(k+1) a pole of the Gamma function in the denominator of (119d) requires that we use Eq.(116) again, leading to I~2=I~1.\tilde{I}_{2}=\tilde{I}_{1}. This can also be found from the linear system (110) by noting that, besides (118), now s(bR+cR)=0s(b_{R}+c_{R})=0.

All of the peculiarities above are taken into account if we simply replace the hypergeometrics by the well-behaved regularized hypergeometric,

I1(aR,bR,cR)\displaystyle I_{1}(a_{R},b_{R},c_{R}) =π(4n2)32wn2F(32,6+n4;3;wn)\displaystyle=\frac{\pi(4-n^{2})}{32}\,w_{n}^{2}\,F(\tfrac{3}{2},\tfrac{6+n}{4};3;w_{n}) (119a)
I2(aR,bR,cR)\displaystyle I_{2}(a_{R},b_{R},c_{R}) =12Γ(6n4)Γ(6+n4)wn2F(32,6+n4;3;wn)\displaystyle=\tfrac{1}{2}\Gamma(\tfrac{6-n}{4})\Gamma(\tfrac{6+n}{4})w_{n}^{2}F(\tfrac{3}{2},\tfrac{6+n}{4};3;w_{n}) (119b)
I~1(aR,bR,cR)\displaystyle\tilde{I}_{1}(a_{R},b_{R},c_{R}) =2πΓ(6+n4)𝐅(n24,12;n+44;1wn)\displaystyle=-2\sqrt{\pi}\,\Gamma(\tfrac{6+n}{4})\,{\bf F}(\tfrac{n-2}{4},-\tfrac{1}{2};\tfrac{n+4}{4};1-w_{n}) (119c)
I~2(aR,bR,cR)\displaystyle\tilde{I}_{2}(a_{R},b_{R},c_{R}) =2π(1wn)n4Γ(6n4)𝐅(n+24,12;4n4;1wn)\displaystyle=-\frac{2\sqrt{\pi}}{(1-w_{n})^{\frac{n}{4}}}\,\Gamma(\tfrac{6-n}{4})\,{\bf F}(-\tfrac{n+2}{4},-\tfrac{1}{2};\tfrac{4-n}{4};1-w_{n}) (119d)

We can now use Eqs.(107) and (99) to write

JR(n)=(n+132n)2[cos(nπ4)I~1(n)I2(n)+I1(n)I~2(n)],{n4k+2kJ_{R}(n)=-\left(\frac{n+1}{32n}\right)^{2}\Big{[}\cos\left(\frac{n\pi}{4}\right)\tilde{I}_{1}(n)I_{2}(n)+I_{1}(n)\tilde{I}_{2}(n)\Big{]},\quad\begin{cases}n\neq 4k+2\\ k\in\mathbb{N}\end{cases} (120)

Before we analyze this result further, let us consider what happens if n=4k+2n=4k+2.


The case n=4k+2n=4k+2

When n=4k+2n=4k+2, a pole of the Gamma function appears in the numerator of (119d), so I(n)I(n) is infinite. We can isolate the divergence, however. First, we list again the four canonical integrals, now in terms of k=n24k=\frac{n-2}{4},

I1(k)\displaystyle I_{1}(k) =32πk(k+1)(2k+1)2(4k+3)4F(32,k+2;3;8(2k+1)(4k+3)2)\displaystyle=-\frac{32\pi k(k+1)(2k+1)^{2}}{(4k+3)^{4}}\,F\left(\tfrac{3}{2},k+2;3;\tfrac{8(2k+1)}{(4k+3)^{2}}\right) (121a)
I2(k)\displaystyle I_{2}(k) =32(2k+1)2(4k+3)4Γ(1k)Γ(k+2)F(32,k+2;3;8(2k+1)(4k+3)2)\displaystyle=\frac{32(2k+1)^{2}}{(4k+3)^{4}}\,\Gamma(1-k)\Gamma(k+2)\,F\left(\tfrac{3}{2},k+2;3;\tfrac{8(2k+1)}{(4k+3)^{2}}\right) (121b)
I~1(k)\displaystyle\tilde{I}_{1}(k) =2πΓ(k+2)𝐅(12,k;k+32;(4k+1)2(4k+3)2)\displaystyle=-2\sqrt{\pi}\,\Gamma(k+2)\,{\bf F}\left(-\tfrac{1}{2},k;k+\tfrac{3}{2};\tfrac{(4k+1)^{2}}{(4k+3)^{2}}\right) (121c)
I~2(k)\displaystyle\tilde{I}_{2}(k) =2π(4k+3)2k+1(4k+1)2k+1Γ(1k)𝐅(12,k1;12k;(4k+1)2(4k+3)2)\displaystyle=-\frac{2\sqrt{\pi}(4k+3)^{2k+1}}{(4k+1)^{2k+1}}\,\Gamma(1-k)\,{\bf F}\left(-\tfrac{1}{2},-k-1;\tfrac{1}{2}-k;\tfrac{(4k+1)^{2}}{(4k+3)^{2}}\right) (121d)

Here we note that in this case we have s(aR)=s(cR)=0s(a_{R})=s(c_{R})=0 besides (118), and the linear system (110) is not valid anymore. This is related to the fact that there is now only one branch point in the canonical integrals, instead of the three branchings of the general case. Eq.(107) is, however, still valid. Moreover, we have I2(k)=I1(k)/sin(πk)I_{2}(k)=-I_{1}(k)/\sin(\pi k). The sine is cancelled in Eq.(107),

I(k)=I1(k)(I~1(k)+I~2(k)).I(k)=-I_{1}(k)\big{(}\tilde{I}_{1}(k)+\tilde{I}_{2}(k)\big{)}. (122)

Now the only singularity comes from the pole of Γ(1k)\Gamma(1-k) in (121d). Making kk+ϵk\to k+\epsilon, we have (cf. bateman1953higher , §1.17 Eq.(11))

Γ(1kϵ)=(1)k1(k1)![1ϵ+ψ(k)+O(ϵ)],\Gamma(1-k-\epsilon)=\frac{(-1)^{k-1}}{(k-1)!}\left[-\frac{1}{\epsilon}+\psi(k)+{\mathrm{O}}(\epsilon)\right], (123)

from which we separate the finite part and the divergence:

I~2(k)\displaystyle\tilde{I}_{2}(k) 1ϵI~2sing(k)+I~2reg(k)\displaystyle\equiv\frac{1}{\epsilon}\tilde{I}_{2}^{sing}(k)+\tilde{I}_{2}^{reg}(k) (124)
I~2sing(k)\displaystyle\tilde{I}_{2}^{sing}(k) =(1)k2π(4k+3)2k+1(4k+1)2k+1(k1)!𝐅(12,k1;12k;(4k+1)2(4k+3)2)\displaystyle=\frac{(-1)^{k}2\sqrt{\pi}(4k+3)^{2k+1}}{(4k+1)^{2k+1}(k-1)!}\,{\bf F}\left(-\tfrac{1}{2},-k-1;\tfrac{1}{2}-k;\tfrac{(4k+1)^{2}}{(4k+3)^{2}}\right) (125)
I~2reg(k)\displaystyle\tilde{I}_{2}^{reg}(k) =(1)k12π(4k+3)2k+1ψ(k)(4k+1)2k+1(k1)!𝐅(12,k1;12k;(4k+1)2(4k+3)2)\displaystyle=\frac{(-1)^{k-1}2\sqrt{\pi}(4k+3)^{2k+1}\psi(k)}{(4k+1)^{2k+1}(k-1)!}\,{\bf F}\left(-\tfrac{1}{2},-k-1;\tfrac{1}{2}-k;\tfrac{(4k+1)^{2}}{(4k+3)^{2}}\right) (126)

where ψ(ζ)\psi(\zeta) is the digamma function. Using (122) and (99), we end up with

JR(k)\displaystyle J_{R}(k) =JRreg(k)+ϵ1JRsing(k)\displaystyle=J_{R}^{reg}(k)+\epsilon^{-1}J_{R}^{sing}(k) (127)
JRreg(k)\displaystyle J_{R}^{reg}(k) =(3+4k128(1+2k))2I1(k)[I~1(k)+I~2reg(k)],\displaystyle=-\left(\frac{3+4k}{128(1+2k)}\right)^{2}\ I_{1}(k)\ \big{[}\tilde{I}_{1}(k)+\tilde{I}_{2}^{reg}(k)\big{]}, (128)
JRsing(k)\displaystyle J_{R}^{sing}(k) =(3+4k128(1+2k))2I1(k)I~2sing(k).\displaystyle=-\left(\frac{3+4k}{128(1+2k)}\right)^{2}\ I_{1}(k)\tilde{I}_{2}^{sing}(k). (129)

which is the final regularized expression for JRJ_{R} when n=4k+2n=4k+2.

Refer to caption
Figure 2: JR(n)J_{R}(n) (and its regularization) for charged Ramond fields.

Comments

We present a unified plot of JR(n)J_{R}(n) for every nn in Fig.2; for n=4k+2n=4k+2, we plot the regularized function JRreg(k)J_{R}^{reg}(k). One can distinguish a peculiar “almost periodicity” of the function, with period 4. We believe that this might be related to some combinatoric relation between the twists of Rn±R^{\pm}_{n} and the twists of the interaction operators appearing in the four-point function. As can be seen from Fig.2, JR(n)J_{R}(n) stabilizes around small, negative values for large nn. As a reference, for k=30k=30 we have

JR(4k+1)0.0215279\displaystyle J_{R}(4k+1)\approx-0.0215279\qquad JRreg(4k+2)0.0036010\displaystyle J_{R}^{reg}(4k+2)\approx-0.0036010
JR(4k+3)0.0354618\displaystyle J_{R}(4k+3)\approx-0.0354618\qquad JR(4k+4)0.0284950\displaystyle J_{R}(4k+4)\approx-0.0284950

Note that an analytic form of JR(n)J_{R}(n) for large nn is very hard to find because it involves taking simultaneous limits of the multiple arguments of the hypergeometric function.

6.3 The integral for bare twists

The function Gσ(x)G_{\sigma}(x) also has the form (96), and

Jσ(n)=d2x|u(x)Gσ(x)|2=[12n(n+1)Cσ]2I(n),\displaystyle J_{\sigma}(n)=\int d^{2}x\;|u^{\prime}(x)G_{\sigma}(x)|^{2}=\left[\tfrac{1}{2}n(n+1)C_{\sigma}\right]^{2}\;I(n), (130)

where I(n)I(n) is a DF integral (100) with exponents

aσ=(n+1)24n,bσ=32,cσ=(n1)24n.a_{\sigma}=\frac{(n+1)^{2}}{4n},\qquad b_{\sigma}=-\frac{3}{2},\qquad c_{\sigma}=-\frac{(n-1)^{2}}{4n}. (131)

The canonical integrals

I1(aσ,bσ,cσ)\displaystyle I_{1}(a_{\sigma},b_{\sigma},c_{\sigma}) =π(n1)22(n+1)2F(32,1+6n+n24n;3;wn)\displaystyle=-\frac{\pi(n-1)^{2}}{2(n+1)^{2}}\,F\left(\tfrac{3}{2},\tfrac{1+6n+n^{2}}{4n};3;w_{n}\right) (132a)
I2(aσ,bσ,cσ)\displaystyle I_{2}(a_{\sigma},b_{\sigma},c_{\sigma}) =8n2(n+1)4Γ(1(n1)24n)Γ(1+(n+1)24n)F(32,1+6n+n24n;3;wn)\displaystyle=\frac{8n^{2}}{(n+1)^{4}}\Gamma\left(1-\tfrac{(n-1)^{2}}{4n}\right)\Gamma\left(1+\tfrac{(n+1)^{2}}{4n}\right)F\left(\tfrac{3}{2},\tfrac{1+6n+n^{2}}{4n};3;w_{n}\right) (132b)
I~1(aσ,bσ,cσ)\displaystyle\tilde{I}_{1}(a_{\sigma},b_{\sigma},c_{\sigma}) =2πΓ(1+(n+1)24n)𝐅(12,(n1)24n;1+4n+n24n;1wn)\displaystyle=-2\sqrt{\pi}\,\Gamma\big{(}1+\tfrac{(n+1)^{2}}{4n}\big{)}\,{\bf F}\left(-\tfrac{1}{2},\tfrac{(n-1)^{2}}{4n};\tfrac{1+4n+n^{2}}{4n};1-w_{n}\right) (132c)
I~2(aσ,bσ,cσ)\displaystyle\tilde{I}_{2}(a_{\sigma},b_{\sigma},c_{\sigma}) =2πΓ(1(n1)24n)(1wn)1+n24n𝐅(12,(1+n)24n;14n+n24n;1wn)\displaystyle=-2\sqrt{\pi}\,\Gamma\big{(}1-\tfrac{(n-1)^{2}}{4n}\big{)}\,(1-w_{n})^{-\frac{1+n^{2}}{4n}}\,{\bf F}\left(-\tfrac{1}{2},-\tfrac{(1+n)^{2}}{4n};-\tfrac{1-4n+n^{2}}{4n};1-w_{n}\right) (132d)

are all well-defined and convergent for all values of nn\in\mathbb{N} — the arguments of the Gamma functions are never a negative integer (for n>1n>1).

We plot the values of

Jσ(n)=(n+132n)2[s(a)I~1(n)I2(n)+s(b)I1(n)I~2(n)]J_{\sigma}(n)=-\left(\frac{n+1}{32n}\right)^{2}\left[s(a)\tilde{I}_{1}(n)I_{2}(n)+s(b)I_{1}(n)\tilde{I}_{2}(n)\right] (133)

in Fig.3. We can see again an approximate periodicity, with period 4, similar to what happens in the Ramond case. We can give the following numerical values for k=30k=30:

Jσ(4k+1)0.0214398\displaystyle J_{\sigma}(4k+1)\approx-0.0214398\qquad Jσ(4k+2)1.11106\displaystyle J_{\sigma}(4k+2)\approx-1.11106
Jσ(4k+3)0.035381\displaystyle J_{\sigma}(4k+3)\approx-0.035381\qquad Jσ(4k+4)0.028456\displaystyle J_{\sigma}(4k+4)\approx-0.028456

to be compared with the corresponding values for JR(n)J_{R}(n) given above. For n=4k+2n=4k+2, Jσ(n)J_{\sigma}(n) grows with nn, instead of stabilizing around a small value; note that these values of nn are also those for which the Ramond integral JR(n)J_{R}(n) diverged, and had to be regularized.

Refer to caption
Figure 3: Numerical result of the integral (130) for twist fields.

6.4 Renormalization of Ramond and twist fields

The renormalized dimension of the Ramond operators is given by Eq.(25). To first order in λ\lambda, it would be proportional do the structure constant of the three-point function

R[n]()O[2](int)(z)R[n]+(0)=0,\langle R^{-}_{[n]}(\infty)O^{(\mathrm{int})}_{[2]}(z)R^{+}_{[n]}(0)\rangle=0, (134)

which vanishes at the free orbifold point because R[n]R^{-}_{[n]} does not appear in the OPE O[2](int)(z,z¯)R[n]+(0)O^{(\mathrm{int})}_{[2]}(z,\bar{z})R^{+}_{[n]}(0). We are thus left with a correction only at order λ2\lambda^{2},

ΔnR(λ)=n2+π2λ2|JR(n)|+,\Delta^{R}_{n}(\lambda)=\frac{n}{2}+\frac{\pi}{2}\lambda^{2}|J_{R}(n)|+\cdots, (135)

where 12n=hnR(0)+h~nR(0)=ΔnR(0)\frac{1}{2}n=h^{R}_{n}(0)+\tilde{h}^{R}_{n}(0)=\Delta^{R}_{n}(0) is the (total) dimension of R[n]±(z,z¯)R^{\pm}_{[n]}(z,\bar{z}) in the free theory. Here it should be understood that JR(4k+2)JRreg(k)J_{R}(4k+2)\equiv J_{R}^{reg}(k). The renormalized fields are therefore given by

R[n]±(ren)(z,z¯)=Λπ2λ2JR(n)R[n]±(z,z¯),R^{\pm(ren)}_{[n]}(z,\bar{z})=\Lambda^{\frac{\pi}{2}\lambda^{2}J_{R}(n)}R^{\pm}_{[n]}(z,\bar{z}), (136)

where Λ1\Lambda\ll 1 is the cutoff appearing in (19). We can also give the renormalization of the structure constant; from Eq.(26),

Rn()O[2](int)(1)Rn+(0)=λJR(n)+\big{\langle}R^{-}_{n}(\infty)O^{(\mathrm{int})}_{[2]}(1)R^{+}_{n}(0)\big{\rangle}=\lambda J_{R}(n)+\cdots (137)

We thus have concluded that for generic n<Nn<N the dimension of the Ramond field slightly increases in the perturbed theory. The case n=2n=2, which is the smallest value of nn for a twisted field,111111We note that the results above do not apply directly to the untwisted Ramond field (n=1n=1). In particular, the DF integral has a different structure, with one less critical point. is special. The function (61) loses the singularity at x=0x=0. It is interesting to see how this arises as a consequence of the permutation structure of the twists: we have seen that the OPE channel with x0x\to 0 results in a field with twist n1n-1; in this case that would be an untwisted field. As a consequence of Eq.(44), this would mean that the 2-cycles of O2(int)O^{(\mathrm{int})}_{2} and R2±R^{\pm}_{2} entering the correlators in this channel are all the same; but such a correlator would involve only 𝐬=2=n{\bf s}=2=n copies instead of n+1n+1, and therefore require a genus-one covering surface. Most importantly, the n=2n=2 twisted Ramond fields do not renormalize — it can be easily checked that

JR(2)=0.J_{R}(2)=0. (138)

For that, it suffices to look at Eqs.(120) and (119), since cos(π2)=0\cos(\frac{\pi}{2})=0 and I1(2)=0I_{1}(2)=0 while the other canonical functions are finite.


The same analysis holds for the bare twists: their dimension in the perturbed theory becomes

Δnσ(λ)=12(n1n)+π2λ2|Jσ(n)|+,\Delta^{\sigma}_{n}(\lambda)=\frac{1}{2}\left(n-\frac{1}{n}\right)+\frac{\pi}{2}\lambda^{2}|J_{\sigma}(n)|+\cdots, (139)

and the renormalized twist operators are

σ[n](ren)(z,z¯)=Λπ2λ2Jσ(n)σ[n](z,z¯).\sigma_{[n]}^{(ren)}(z,\bar{z})=\Lambda^{\frac{\pi}{2}\lambda^{2}J_{\sigma}(n)}\sigma_{[n]}(z,\bar{z}). (140)

The structure constant

σ[n]()O[2](int)(1)σ[n](0)=λJσ(n)+\langle\sigma_{[n]}(\infty)O^{(\mathrm{int})}_{[2]}(1)\sigma_{[n]}(0)\big{\rangle}=\lambda J_{\sigma}(n)+\cdots (141)

which also vanishes in the free theory, acquires a non-vanishing value at first-order.


The regularization of the divergent integral JJ described in this section gives well-defined, finite two-point functions in the deformed theory, to second order in λ\lambda. Here we have considered the renormalization of bare twists and Ramond fields, but the method is more general, and can be applied to all sectors of the SCFT2. Our procedure relied on the fact that JJ can be reduced to a Dotsenko-Fateev integral for the functions GR(x)G_{R}(x) and Gσ(x)G_{\sigma}(x). This, in turn, relied on the structure of these functions, which had the form (96). It is not hard to check that, for any primary twisted 𝒪[n]\mathscr{O}_{[n]} that we insert in the general correlation function (27), the corresponding G(x)G(x) always has the form (96), including the specific relations between the pairs of exponents α1,α3\alpha_{1},\alpha_{3} and α2,α4\alpha_{2},\alpha_{4}.

To see that this is true, one can reverse-engineer the reasoning developed in Sect.5. Given an operator 𝒪n\mathscr{O}_{n} consider the correlator

G(u,u¯)=𝒪[n]()O[2](int)(1)O[2](int)(u,u¯)𝒪[n](0).G(u,\bar{u})=\big{\langle}\mathscr{O}_{[n]}^{\dagger}(\infty)O^{(\mathrm{int})}_{[2]}(1)O^{(\mathrm{int})}_{[2]}(u,\bar{u})\mathscr{O}_{[n]}(0)\big{\rangle}.

This function must be singular in the short-distance limit u0u\to 0, and consistent with the OPE rule (65). The associated function G(x)G(x) must therefore be singular when xx goes to one of the values x𝔞0x^{0}_{\mathfrak{a}} for u0u\to 0, or x𝔞x^{\infty}_{\mathfrak{a}} for uu\to\infty, where x𝔞0x^{0}_{\mathfrak{a}}, x𝔞x^{\infty}_{\mathfrak{a}} are the channels in the limits u0u\to 0 and uu\to\infty, respectively; see App.B. This fixes the numerator of

G(x)=C(xx10)α1(xx2)α2(xx20)α3(xx1)α4(xx21)α5G(x)=C\frac{(x-x^{0}_{\mathfrak{1}})^{\alpha_{1}}(x-x^{\infty}_{\mathfrak{2}})^{\alpha_{2}}(x-x^{0}_{\mathfrak{2}})^{\alpha_{3}}(x-x^{\infty}_{\mathfrak{1}})^{\alpha_{4}}}{(x-x^{1}_{\mathfrak{2}})^{\alpha_{5}}} (142)

while the denominator is fixed similarly by the channels in u1u\to 1. But (142) is just another way to write (96). Note that this argument only makes use of the properties of the function u(x)u(x) and its inverses, i.e. only on the structure of the twists in the correlator — not on the specifics of 𝒪n\mathscr{O}_{n} nor, even, on the properties of O2(int)O^{(\mathrm{int})}_{2}. Thus 𝒪[n]\mathscr{O}_{[n]} can be, say, a primary NS field, an R-charged or R-neutral Ramond ground state, or a bare twist field; also, we can replace O[2](int)O^{(\mathrm{int})}_{[2]} by, say, the simplest chiral NS primaries O[2](p,q)O^{(p,q)}_{[2]} (defined e.g. in Skenderis:2008qn ).

Having proved that G(x)G(x) must have the structure (142), it remains for us to show that the exponents satisfy the two relations in (96). This is also a consequence of the OPEs. Take the channel x10x^{0}_{\mathfrak{1}} in the limit u0u\to 0. We have the OPE O2(int)𝒪n𝒳n1O^{(\mathrm{int})}_{2}\mathscr{O}_{n}\sim\mathscr{X}_{n-1} for some operator of twist n1n-1, whose dimension is fixed by the power of uu appearing in G(x10(u))G(x^{0}_{\mathfrak{1}}(u)). Since G(x10(u))[x10(u)]α1G(x^{0}_{\mathfrak{1}}(u))\sim[x^{0}_{\mathfrak{1}}(u)]^{\alpha_{1}}, using (165) we have G(x10(u))uα1n1G(x^{0}_{\mathfrak{1}}(u))\sim u^{\frac{\alpha_{1}}{n-1}}, hence the holomorphic dimension of 𝒳n1\mathscr{X}_{n-1} is

hn1𝒳=α1n1+hn𝒪+h2O(int).h^{\mathscr{X}}_{n-1}=\frac{\alpha_{1}}{n-1}+h^{\mathscr{O}}_{n}+h^{O^{(\mathrm{int})}}_{2}. (143)

Now, in the limit uu\to\infty, we will have the OPE O2(int)𝒪n𝒳n1O^{(\mathrm{int})}_{2}\mathscr{O}_{n}^{\dagger}\sim\mathscr{X}_{n-1}^{\dagger}. Using (168), we now have G(x1(u))uα4n1G(x^{\infty}_{\mathfrak{1}}(u))\sim u^{\frac{\alpha_{4}}{n-1}}, and the dimension of 𝒳n1\mathscr{X}^{\dagger}_{n-1} is

hn1𝒳=α4n1+hn𝒪h2O(int),h^{\mathscr{X}^{\dagger}}_{n-1}=\frac{\alpha_{4}}{n-1}+h^{\mathscr{O}}_{n}-h^{O^{(\mathrm{int})}}_{2}, (144)

with a minus sign in front of h2O(int)h^{O^{(\mathrm{int})}}_{2} because we must conjugate O2(int)O^{(\mathrm{int})}_{2} to \infty. But 𝒳m\mathscr{X}_{m} and 𝒳m\mathscr{X}^{\dagger}_{m} have the same dimension, so subtracting Eqs.(143) and (144) we find

α1α4n1=2h2O(int)\frac{\alpha_{1}-\alpha_{4}}{n-1}=2h^{O^{(\mathrm{int})}}_{2}

which, since h2O(int)=1h^{O^{(\mathrm{int})}}_{2}=1, gives the first relation in (96). The second relation, between α1\alpha_{1} and α3\alpha_{3}, is found similarly in the channels x20x^{0}_{\mathfrak{2}} and x2x^{\infty}_{\mathfrak{2}}, completing the proof that (96) holds in general.

Thus we have shown that, for any primary twisted filed 𝒪[n]\mathscr{O}_{[n]}, we can always reduce J𝒪J_{\mathscr{O}} to a Dotsenko-Fateev integral, for some set of parameters a,b,ca,b,c. Then, we can apply our regularization procedure and subsequent renormalization of the two-point function 𝒪[n]𝒪[n]\langle\mathscr{O}_{[n]}^{\dagger}\mathscr{O}_{[n]}\rangleif that is necessary. A very important example of fields for which there is no renormalization is the class of BPS-protected NS chiral twisted fields. Explicit computation of their non-renormalization was given in Pakman:2009mi , for On(0,0)O^{(0,0)}_{n}, the lowest-weight operator in the nn-twisted sector of the NS chiral ring Jevicki:1998bm , with hnNS=n12=j3h^{\mathrm{NS}}_{n}=\frac{n-1}{2}=j^{3}. (The descendants of O2(0,0)O^{(0,0)}_{2} give the deformation operator O2(int)O^{(\mathrm{int})}_{2}.) The four-point function GO(x)G_{O}(x) was found by the same method of Sect.4, see Eq.(D.6) of Ref.Pakman:2009mi . It has the form (96), and gives rise to a Dotsenko-Fateev integral with exponents aO=1a_{O}=1, bO=32b_{O}=-\tfrac{3}{2}, cO=0c_{O}=0. Since cO=0c_{O}=0, the integral (100) simplifies, and can be computed directly in terms of Gamma functions, as done in App.D of Ref.Pakman:2009mi , without the need to resorting to the hypergeometric regularization machinery. Nevertheless, it is interesting to confirm that our formulae do give the same result, i.e. JO=0J_{O}=0. Inserting aOa_{O}, bOb_{O} and cOc_{O} into our canonical functions (112)-(113), we find

I1=0,I2=42(wn2)1wn,I~1=4,I~2=2(wn2)1wnI_{1}=0,\quad I_{2}=-4-\frac{2\left(w_{n}-2\right)}{\sqrt{1-w_{n}}},\quad\tilde{I}_{1}=-4,\quad\tilde{I}_{2}=\frac{2(w_{n}-2)}{\sqrt{1-w_{n}}}

and s(aO)=0s(a_{O})=0, so

JO(n)=[12n(n+1)CO]2(s(aO)I~1(n)I2(n)+s(bO)I1(n)I~2(n))=0J_{O}(n)=-\left[\tfrac{1}{2}n(n+1)C_{O}\right]^{2}\Big{(}s(a_{O})\tilde{I}_{1}(n)I_{2}(n)+s(b_{O})I_{1}(n)\tilde{I}_{2}(n)\Big{)}=0 (145)

as expected.

Let us point out that our regularization and renormalization procedure is even more general. It can be extended almost intactly for the analysis of two-point functions of operators with a more complicated twist structure. In Ref.Lima:2020nnx we have studied the double-cycle composite Ramond fields R[n]±R[m]±(z,z¯)R^{\pm}_{[n]}R^{\pm}_{[m]}(z,\bar{z}). In this case, the covering map is more complicated, and, correspondingly, so is the form of G(x)G(x) which generalizes (96); but just as explained above, there are relations between exponents which allow a transformation of J(n,m)J(n,m) into a Dotsenko-Fateev integral, and then everything follows as in here.

6.5 On spectral flow

The spectral flow automorphism of the 𝒩=4\mathcal{N}=4 super-algebra Schwimmer:1986mf acts on the Virasoro and R-current modes as

LL=LJ3ξ+c24ξ2δ,0J3J3=J3c12ξδ,0\displaystyle\begin{split}L_{\ell}&\mapsto L_{\ell}^{\prime}=L_{\ell}-J^{3}_{\ell}\xi+\frac{c}{24}\xi^{2}\delta_{\ell,0}\\ J^{3}_{\ell}&\mapsto J^{\prime 3}_{\ell}=J^{3}_{\ell}-\frac{c}{12}\,\xi\,\delta_{\ell,0}\end{split} (146)

where ξ\xi is the spectral flow parameter. Hence an operator with conformal weight hh and R-charge j3j^{3} is mapped to an operator with

h=hj3ξ+c24ξ2,j3=j3c12ξ,\displaystyle\begin{split}h^{\prime}&=h-j^{3}\,\xi+\frac{c}{24}\xi^{2}\,,\qquad j^{\prime 3}=j^{3}-\frac{c}{12}\xi,\end{split} (147)

and NS (anti-)chiral fields flow to Ramond ground states:

hNS=±j3h=124cforξ=±1.h^{\mathrm{NS}}=\pm j^{3}\quad\mapsto\quad h^{\prime}=\tfrac{1}{24}c\quad\text{for}\quad\xi=\pm 1. (148)

Our renormalized fields R[n]±R^{\pm}_{[n]}, which are Ramond ground states of the nn-wound string, have conformal weight

hnR=14n<14N,h^{\mathrm{R}}_{n}=\tfrac{1}{4}n<\tfrac{1}{4}N, (149)

the bound n<Nn<N being due to our calculation at order N1N^{-1}; the field with n=Nn=N scales as N2N^{-2} for large NN, and requires a genus-one covering map. In the free theory, in the presence of a nn-twist, it is possible to consider the ZnZ_{n} orbifold 𝒩=4\mathcal{N}=4 SCFT with c=6nc=6n, whose conserved currents are defined by adding the nn copies entering the twist. For example, taking the cycle to be (1,,n)(1,\cdots,n), the nn-twisted CFT has stress tensor and Virasoro modes121212In the presence of σ(n)\sigma_{(n)} we can also define the usual fractional modes LknL_{\frac{k}{n}}, Jkn3J^{3}_{\frac{k}{n}}, etc., as well as a ‘fractional spectral flow’ which is an automorphism of the fractional algebra deBeer:2019ioe .

T(z)=I=1nTI(z),Lkσ(1,,n)(0)=dz2πiz1+kI=1nTI(z)σ(1,,n)(0).T(z)=\sum_{I=1}^{n}T_{I}(z)\,,\qquad L_{k}\;\sigma_{(1,\cdots,n)}(0)=\oint\frac{dz}{2\pi i}z^{1+k}\sum_{I=1}^{n}T_{I}(z)\sigma_{(1,\cdots,n)}(0). (150)

The modes LkL_{k}, with kk\in\mathbb{Z}, are well defined: the twist shuffles the terms in the summation over II, but the summation itself is preserved. Spectral flow of this nn-twisted algebra by ξ=1\xi=1, when applied to the NS chiral field O(1,,n)(0,0)O^{(0,0)}_{(1,\cdots,n)} with hnNS=n12=jn3h^{\mathrm{NS}}_{n}=\frac{n-1}{2}=j^{3}_{n}, gives the Ramond field R(1,,n)R^{-}_{(1,\cdots,n)}, with h=14nh^{\prime}=\frac{1}{4}n and j3=12j^{\prime 3}=-\frac{1}{2}. Starting with the anti-chiral NS and flowing by ξ=1\xi=-1, we get R(1,,n)+R^{+}_{(1,\cdots,n)}, etc.

As shown by Eq.(145), the dimension of the field O(1,,n)(0,0)O^{(0,0)}_{(1,\cdots,n)} is protected in the deformed theory, although R(1,,n)±R^{\pm}_{(1,\cdots,n)} is renormalized. How is this to be reconciled with the spectral flow between them? The answer is that the symmetry algebra of the nn-twisted CFT, i.e. the 𝒩=(4,4)\mathcal{N}=(4,4) SCFT with central charge c=6n<6Nc=6n<6N, is not preserved after the deformation by the interaction O[2](int)O^{(\mathrm{int})}_{[2]}. The basic reason for this is that the twist can join two strings into a longer string.

Let us discuss this in more detail. The deformed SCFT has deformed charges T(λ)(z)T^{(\lambda)}(z), J(λ)a(z)J^{(\lambda)a}(z), G(λ)αA(z)G^{(\lambda)\alpha A}(z), which must close under an operator algebra. If this algebra has an automorphism, then we can define a spectral flow between the deformed states. The deformed charges are difficult to describe explicitly. In particular, obtaining the stress-tensor T(λ)(z)T^{(\lambda)}(z) is subtle, as explained in Guo:2019pzk , since one cannot naïvely make a variation of the action (16). Instead, one can use the prescription of SenSEN1990551 to obtain the Virasoro modes by the following action on a field Φ\Phi,

Lk(λ)Φ(0)=(Lk+λδLk)Φ0=LkΦ(0)+λ|z|=ϵdz¯zk+1O[2](int)(z,z¯)Φ(0).L_{k}^{(\lambda)}\Phi(0)=(L_{k}+\lambda\delta L_{k})\Phi_{0}=L_{k}\Phi(0)+\lambda\underset{|z|=\epsilon}{\oint}\!d\bar{z}\ z^{k+1}\,O^{(\mathrm{int})}_{[2]}(z,\bar{z})\Phi(0). (151)

See also Guo:2019pzk , and Campbell:1990dz for a discussion of this formula. As shown in SEN1990551 , the modes Lk(λ)L^{(\lambda)}_{k} satisfy the Virasoro algebra with the same central charge as the unperturbed algebra of the LkL_{k}, which are derived from the unperturbed tensor T(z)T(z). Such preservation of the Virasoro algebra after the deformation (151) requires computing [Lk,δLl][L_{k},\delta L_{l}], which in turn includes computing the integral

λ𝑑zz around wzl+1T(z)O[2](int)(w,w¯).\underset{\text{$z$ around $w$}}{\lambda\ \oint\!dz}z^{l+1}\,T(z)O^{(\mathrm{int})}_{[2]}(w,\bar{w}). (152)

Now, O[2](int)O^{(\mathrm{int})}_{[2]} is made out of the sum over the conjugacy class of Z2Z_{2} cycles in SNS_{N}, hence it involves all the NN copies of the orbifold. Therefore the integral above is only defined if T(z)T(z) also includes all the NN copies of the fields. For example, if we take the nn-twisted algebra made by the modes (150), when going around the twist σ(n,n+1)\sigma_{(n,n+1)} the integral would not be defined. In other words, the Virasoro algebra with c=6nc=6n is inconsistent with the deformation (16) for n<Nn<N.

Let us also stress another important detail of our computation of the corrected dimension: since we perform the integral JRJ_{R} by changing coordinates to the covering surface, we automatically include all the conjugacy classes of the permutations inside the four-point function, which are taken into account by the very nature of the covering map, as we have extensively discussed in Sect.5. Thus we are truly computing the renormalized fields in the theory deformed by O[2](int)O^{(\mathrm{int})}_{[2]}, rather than by a non-SNS_{N}-invariant operator such as, say, O(1,2)(int)O^{(\mathrm{int})}_{(1,2)}.

Thus spectral flow between R[n]±R^{\pm}_{[n]} and O[n](0,0)O^{(0,0)}_{[n]} is not preserved after the deformation by O[2](int)O^{(\mathrm{int})}_{[2]}, unless n=Nn=N. Spectral flow of the full orbifold theory, with c=6Nc=6N is, however, (expected to be) preserved. But, in the full orbifold theory, R[n]±R^{\pm}_{[n]} is not a “true” Ramond ground state — it is a mix of a nn-twisted string in a Ramond ground state, with NnN-n untwisted strings in the NS ground state. Thus one would not expect R[n]±R^{\pm}_{[n]} to be BPS protected. Explicitly, taking ξ=1\xi=1 and with c=6Nc=6N, the field R[n]R^{-}_{[n]} with hnR=14nh^{\mathrm{R}}_{n}=\frac{1}{4}n and j3=12j^{3}=-\frac{1}{2} flows to a state with

h=n+N24,j3=N12h^{\prime}=\frac{n+N-2}{4},\qquad j^{\prime 3}=\frac{N-1}{2} (153)

which is only chiral for n=Nn=N. Conversely, the NS chiral O[n](0,0)O^{(0,0)}_{[n]} flows to a state with

h=14N,j3=Nn+12h^{\prime}=\tfrac{1}{4}N,\qquad j^{\prime 3}=-\frac{N-n+1}{2} (154)

which is a “true” Ramond ground state made by composing twisted and untwisted R-charged and R-neutral Ramond fields; when n=Nn=N, this field is simply R[N]R^{-}_{[N]}.

Note that this latter state with twist NN is indeed protected, as far as our computation is concerned, since, as shown in Eq.(48), at leading order in 1/N1/N, the four-point function involving the R[N]±R^{\pm}_{[N]} vanishes, and there is no correction to their dimensions. For the single-cycle fields considered here, this protection is rather trivial, being due simply to the large-NN approximation. But we have shown in Lima:2020nnx that the protection is again observed in composite fields R[m1]αR[m2]β(z,z¯)R^{\alpha}_{[m_{1}]}R^{\beta}_{[m_{2}]}(z,\bar{z}) with dimension hR=14Nh^{\mathrm{R}}=\frac{1}{4}N, for which m1+m2=Nm_{1}+m_{2}=N. In this case, there is a genus-zero contribution to the four-point function, and protection comes from a non-trivial DF integral being zero. Because of spectral flow, we expect that such results generalize for any composite Ramond field whenever the sum of the composing twists add to NN.

7 Discussion

The investigation of the twisted Ramond sector of marginally-deformed D1-D5 SCFT2 presented in this paper is based on the explicit construction of the large-NN limit of the four-point function (50) of two R-charged Ramond fields and two scalar modulus operators O[2](int)(z,z¯)=ϵABG12AG~12˙BO[2](0,0)(z,z¯)O^{(\mathrm{int})}_{[2]}(z,\bar{z})=\epsilon_{AB}G^{-A}_{-\frac{1}{2}}\tilde{G}^{\dot{-}B}_{-\frac{1}{2}}O^{(0,0)}_{[2]}(z,\bar{z}). We have found that R[n]αR^{\alpha}_{[n]} undergoes renormalization for all twists 2<n<N2<n<N, and is protected for the maximal and minimal twist values, n=Nn=N and n=2n=2. The fields σ[n]\sigma_{[n]} also renormalize for n<Nn<N. In fact, these four-point functions provide dynamical information about both theories: the “free-orbifold point” SCFT2, and its marginal deformation, at second order in λ\lambda. In what follows, we will briefly address a some open problems whose solutions can eventually be reached by adapting the methods developed in the present paper.


More on the properties of non-BPS fields. The four-point functions that we have calculated can be used not only for accessing the deformed SCFT2, but also to give a more complete description of the free orbifold itself. For example, the OPE data we have extracted from short-distance limits reveal important features of the Ramond sector of the free SCFT2, such as the conformal weights, R-charges and a few structure constants of the non-BPS twisted Ramond operators Yn±1±Y_{n\pm 1}^{\pm} given by Eq.(95). Their four-point functions with the deformation operator,

Y[m]()O[2](int)(1)O[2](int)(u,u¯)Y[m]+(0),\big{\langle}Y^{-}_{[m]}(\infty)O^{(\mathrm{int})}_{[2]}(1)O^{(\mathrm{int})}_{[2]}(u,\bar{u})Y^{+}_{[m]}(0)\big{\rangle},

can be explicitly constructed by the same covering map and the same methods used here. Computation of this function would provide new relevant CFT data: apart from the corrections to the canonical conformal dimensions (84), it also contains, in the corresponding OPE limits, all the super-conformal properties of the next members of the family of the non-BPS twisted Ramond fields.

Relevant information about fuzzball microstates can be extracted from four-point functions similar to (2), but with the deformation operators replaced by NS chiral fields O[2](p,p)O_{[2]}^{(p,p)}, with dimensions Δ2(0,0)=1\Delta^{(0,0)}_{2}=1 and Δ2(1,1)=2\Delta^{(1,1)}_{2}=2, viz.

R[n]()O[2](p,p)(1)O[2](p,p)(u,u¯)R[m]+(0).\big{\langle}R^{-}_{[n]}(\infty)O^{(p,p)}_{[2]}(1)O^{(p,p)}_{[2]}(u,\bar{u})R^{+}_{[m]}(0)\big{\rangle}. (155)

Computing these functions by the methods of Sect.4 is actually easier than computing (2). Their short-distance limits contain the CFT data — the conformal dimensions and structure constants — about the non-BPS fields Xn±1(p,p)±X^{(p,p)\pm}_{n\pm 1} appearing in the OPEs O[2](p,p)Rn±O^{(p,p)}_{[2]}R_{n}^{\pm}.


R-neutral Ramond ground states. Here we have focused on the R-charged SU(2)L,R doublets R[n]αR^{\alpha}_{[n]}. The remaining Ramond fields, that are neutral under R-symmetry and form a doublet of SU(2)2, have only been mentioned in passing. The renormalization of such fields with a single cycle, R[n]A˙R^{\dot{A}}_{[n]}, can be studied with the same methods of the present paper. One must compute their four-point function with O[2](int)O^{(\mathrm{int})}_{[2]} by lifting to covering space with the corresponding R-neutral spin fields SA˙S^{\dot{A}}, etc. In practice, the actual computation of the four-point function is more complicated then the one presented in §4.2, because some simplifying cancelations only occur for the R-charged fields. Once the function is found, however, all the methodology developed here for exploration of operator algebras via short-distance limits, as well as the renormalization scheme, can be applied. We have presented these results elsewhere Lima:2020urq .


Lifted vs. protected states. Supergravity solutions of fuzzballs correspond to states in the CFT where all component strings are in a Ramond ground state, i.e.  k(R[nk]αk)k\prod_{k}(R^{\alpha_{k}}_{[n_{k}]})^{k}, with kknk=N\sum_{k}kn_{k}=N. If there is only one nn-wound string, then this state becomes (R[1]β)NnR[n]α(R^{\beta}_{[1]})^{N-n}R^{\alpha}_{[n]}, where the untwisted Ramond field R[1]βR^{\beta}_{[1]} is a (symmetrized) spin field. The fields that we have found to be renormalized are made by putting the nn-wound string is in a Ramond ground state, while all the other untwisted strings are in the NS vacuum, i.e. explicitly R[n]α(𝟙)NnR[n]αR^{\alpha}_{[n]}\cong({\mathds{1}})^{N-n}R^{\alpha}_{[n]}. The fact that one can define a spectral flow of the c=6nc=6n super-conformal algebra in the nn-wound string at the free orbifold point may suggest that both (𝟙)NnR[n]α({\mathds{1}})^{N-n}R^{\alpha}_{[n]} and (R[1]β)NnR[n]α(R^{\beta}_{[1]})^{N-n}R^{\alpha}_{[n]} are protected. As we have shown here, this is not true.

The single-cycle Ramond field with maximal twist n=Nn=N, which is a “pure” Ramond ground state in the full orbifold theory, is, indeed, protected, as far as our analysis goes: the four-point function for this field scales as 1/N21/N^{2} and must be computed with a genus-one covering surface. The interesting feature that our calculation highlights is that this protection depends crucially on the combinatorics involved in the combinations of the permutation cycles of the twisted fields R[n]αR^{\alpha}_{[n]} and O[2](int)O^{(\mathrm{int})}_{[2]}. The protection is due not to the fact that the four-point function or its integral vanish — in contrast to what happens in the case of minimal twist n=2n=2, where the behavior of the four-point functions changes drastically and JR(2)J_{R}(2) vanishes, neither the function z(t)z(t), the map u(x)u(x), the functions GR(x)G_{R}(x), nor the integrals JR(n)J_{R}(n), none of them is able to distinguish between 2<n<N2<n<N or n=Nn=N. What separates the maximal twist is the analysis of Eq.(44), which dictates the NN-dependence of GR(u,u¯)G_{R}(u,\bar{u}), and implies that any function obtained from the genus-zero covering map must correspond to a permutation of SNS_{N} such that n<Nn<N. These features become starker when we consider the composite field (R[n1]α1R[n2]α2)(𝟙)Nn1n2R[n1]α1R[n2]α2(R^{\alpha_{1}}_{[n_{1}]}R^{\alpha_{2}}_{[n_{2}]})\cong({\mathds{1}})^{N-n_{1}-n_{2}}R^{\alpha_{1}}_{[n_{1}]}R^{\alpha_{2}}_{[n_{2}]}, which was done in Lima:2020nnx . When n1+n2<Nn_{1}+n_{2}<N, the solutions to the equation equivalent to Eq.(44) allow for factorizations of

(R[n1]α1R[n2]α2)(,¯)O[2](int)(1,1¯)O[2](int)(u,u¯)(R[n1]α1R[n2]α2)(0,0¯)\big{\langle}(R^{\alpha_{1}}_{[n_{1}]}R^{\alpha_{2}}_{[n_{2}]})^{\dagger}(\infty,\bar{\infty})O^{(\mathrm{int})}_{[2]}(1,\bar{1})O^{(\mathrm{int})}_{[2]}(u,\bar{u})(R^{\alpha_{1}}_{[n_{1}]}R^{\alpha_{2}}_{[n_{2}]})(0,\bar{0})\big{\rangle}

into four-point functions involving only one of the single-cycle fields; then the composite field renormalizes as a corollary of our present results. When n1+n2=Nn_{1}+n_{2}=N, and the field becomes a “pure” Ramond ground state with h=124corbh=\frac{1}{24}c_{orb}, there is no factorization. Now, the covering surface for this completely connected function has genus zero, and we can calculate the four-point function and its integral explicitly, in contrast to what happened here for the field R[N]αR^{\alpha}_{[N]}. The four-point function we find in Lima:2020nnx is non-trivial, and reveals conformal data and OPE fusion rules. Its integral, corresponding to JRJ_{R}, is also non-trivial but it does vanish after we apply our regularization procedure and the DF construction: thus we see explicitly that the family of pure Ramond fields (R[n1]α1R[n2]α2)(R^{\alpha_{1}}_{[n_{1}]}R^{\alpha_{2}}_{[n_{2}]}) with n1+n2=Nn_{1}+n_{2}=N is again protected.

In closing, one cannot avoid the question of what are (if any) the bulk holographic images of the renormalized R[n]αR^{\alpha}_{[n]} fields, with their continuous, λ\lambda-dependent conformal dimensions. The answer remains to be discovered, and there are indications that tools necessary for this end include the description of the symmetry algebra of the deformed SCFT2 and its unitary representations.

Acknowledgements.
The work of M.S. is partially supported by the Bulgarian NSF grant KP-06-H28/5 and that of M.S. and G.S. by the Bulgarian NSF grant KP-06-H38/11. M.S. is grateful for the kind hospitality of the Federal University of Espírito Santo, Vitória, Brazil, where part of his work was done.

Appendix A Conventions for the 𝒩=(4,4)\mathcal{N}=(4,4) SCFT

In the 𝒩=(4,4)\mathcal{N}=(4,4) superalgebra, the R-currents Ja(z)J^{a}(z), J~a(z¯)\tilde{J}^{a}(\bar{z}), and the supercurrents GαA(z)G^{\alpha A}(z), G~α˙A˙(z¯)\tilde{G}^{\dot{\alpha}\dot{A}}(\bar{z}) have indices in SU(2) groups as follows: a=1,2,3a=1,2,3 and a˙=1˙,2˙,3˙\dot{a}=\dot{1},\dot{2},\dot{3} transform as a triplets of SU(2)L and SU(2)R, respectively; α=+,\alpha=+,- and α˙=+˙,˙\dot{\alpha}=\dot{+},\dot{-} transform as a doublets of SU(2)L and SU(2)R, respectively; indices A=1,2A=1,2 and A˙=1˙,2˙\dot{A}=\dot{1},\dot{2} transform as doublets of SU(2)1 and SU(2)2, respectively.

The SCFT can be realized in terms of four real bosons Xi(z,z¯)X_{i}(z,\bar{z}), four real holomorphic fermions ψi(z)\psi_{i}(z) and four real anti-holomorphic fermions ψ~i(z¯)\tilde{\psi}_{i}(\bar{z}), with i=1,,4i=1,\cdots,4. They are related to the complex fields XA˙A(z,z¯)X^{\dot{A}A}(z,\bar{z}), ψαA˙(z)\psi^{\alpha\dot{A}}(z) and ψ~α˙A˙(z¯)\tilde{\psi}^{\dot{\alpha}\dot{A}}(\bar{z}) by

XA˙A=12Xi[σi]A˙A=12[X3+iX4X1iX2X1+iX2X3+iX4],\displaystyle X_{\dot{A}A}=\tfrac{1}{\surd 2}X_{i}[\sigma^{i}]_{\dot{A}A}=\frac{1}{\surd 2}\begin{bmatrix}X_{3}+iX_{4}&X_{1}-iX_{2}\\ X_{1}+iX_{2}&-X_{3}+iX_{4}\end{bmatrix}, (156)
ψα1˙=[ψ+1˙ψ1˙]=12[ψ1+iψ2ψ3+iψ4],ψα2˙=[ψ+2˙ψ2˙]=12[ψ3iψ4ψ1+iψ2].\displaystyle\psi^{\alpha\dot{1}}=\begin{bmatrix}\psi^{+\dot{1}}\\ \psi^{-\dot{1}}\end{bmatrix}=\frac{1}{\surd 2}\begin{bmatrix}\psi_{1}+i\psi_{2}\\ \psi_{3}+i\psi_{4}\end{bmatrix},\quad\psi^{\alpha\dot{2}}=\begin{bmatrix}\psi^{+\dot{2}}\\ \psi^{-\dot{2}}\end{bmatrix}=\frac{1}{\surd 2}\begin{bmatrix}\psi_{3}-i\psi_{4}\\ -\psi_{1}+i\psi_{2}\end{bmatrix}. (157)

There are analogous constructions for the right-moving sector. The Levi-Civita symbol always has the structure ϵ12=+1\epsilon^{12}=+1. Pauli matrices are defined such that σ3=Diag(1,1)\sigma^{3}=\mathrm{Diag}(1,-1). The “Pauli vector” σi=(σ1,σ2,σ3,σ4)\sigma^{i}=(\sigma^{1},\sigma^{2},\sigma^{3},\sigma^{4}) and its conjugate σ¯i\bar{\sigma}^{i} have components (we work in Euclidean space) σa=σ¯a\sigma^{a}=-\bar{\sigma}^{a} σ4=i𝟙2×2=σ¯4\sigma^{4}=i\mathds{1}_{2\times 2}=\bar{\sigma}^{4}.

The reality condition of XiX_{i} and ψi\psi_{i} implies that

XA˙AϵA˙B˙ϵABXB˙B,(XA˙A)=XA˙A,(ψαA˙)=ψαA˙.X^{\dot{A}A}\equiv-\epsilon^{\dot{A}\dot{B}}\epsilon^{AB}X_{\dot{B}B},\quad(X_{\dot{A}A})^{\dagger}=X^{\dot{A}A}\ ,\quad(\psi^{\alpha\dot{A}})^{\dagger}=\psi_{\alpha\dot{A}}. (158)

Two-point functions are

XA˙A(z)XB˙B(z)\displaystyle\langle\partial X^{\dot{A}A}(z)\partial X^{\dot{B}B}(z^{\prime})\rangle =2ϵA˙B˙ϵAB(zz)2\displaystyle=\frac{2\epsilon^{\dot{A}\dot{B}}\epsilon^{AB}}{(z-z^{\prime})^{2}} (159)
ψαA˙(z)ψβB˙(z)\displaystyle\langle\psi^{\alpha\dot{A}}(z)\psi^{\beta\dot{B}}(z^{\prime})\rangle =ϵαβϵA˙B˙zz\displaystyle=-\frac{\epsilon^{\alpha\beta}\epsilon^{\dot{A}\dot{B}}}{z-z^{\prime}} (160)
ϕr(z)ϕs(z)\displaystyle\langle\partial\phi_{r}(z)\partial\phi_{s}(z^{\prime})\rangle =δrs(zz)2\displaystyle=-\frac{\delta_{rs}}{(z-z^{\prime})^{2}} (161)

where the last equation is for the bosonized fermions (4). The non-vanishing bosonic two-point functions are between a current XA˙A\partial X^{\dot{A}A} and its complex conjugate; explicitly,

X1˙1(z)(X1˙1)(z)=2(zz)2,X1˙2(z)(X1˙2)(z)=2(zz)2,\langle\partial X^{\dot{1}1}(z)(\partial X^{\dot{1}1})^{\dagger}(z^{\prime})\rangle=\frac{2}{(z-z^{\prime})^{2}},\quad\langle\partial X^{\dot{1}2}(z)(\partial X^{\dot{1}2})^{\dagger}(z^{\prime})\rangle=-\frac{2}{(z-z^{\prime})^{2}}, (162)

as can be checked from (159) using the reality conditions (158).

Appendix B Asymptotics to OPEs

In calculating OPEs, we need to know the inverse of (32) near the base-sphere points u=0,,1u_{*}=0,\infty,1. When u=0u_{*}=0, the roots of Eq.(43) are obvious: x=0x=0 (with multiplicity n1n-1) and x=nx=-n (with multiplicity n+1n+1). Going back to (32), we find the form of u(x)u(x) in these two limits,

u(x)nn+1(1n)n1xn1,u(x)nn1(n1)n+1(x+n)n+1\displaystyle u(x)\approx\frac{n^{n+1}}{(1-n)^{n-1}}x^{n-1},\qquad u(x)\approx\frac{n^{n-1}}{(-n-1)^{n+1}}(x+n)^{n+1}

so inverting we get the two functions

(u0)[x0,x10(u)((1n)n1nn+1u)1n1xn,x20(u)n+((n1)n+1nn1u)1n+1(u\to 0)\quad\left[\begin{array}[]{@{}l@{\quad}l@{}}x\to 0,&x^{0}_{\mathfrak{1}}(u)\approx\left(\frac{(1-n)^{n-1}}{n^{n+1}}u\right)^{\frac{1}{n-1}}\\ x\to-n,&x^{0}_{\mathfrak{2}}(u)\approx-n+\left(\frac{(-n-1)^{n+1}}{n^{n-1}}u\right)^{\frac{1}{n+1}}\end{array}\right. (165)

Taking u=u_{*}=\infty, Eq.(43) reduces to (x1)n+1(x+n1)n1=0(x-1)^{n+1}(x+n-1)^{n-1}=0, with roots x=1x=1 and x=1nx=1-n. The function u(x)u(x) behaves in these limits as

u(x)(1+n)n+1nn11(x1)n+1,u(x)(n1)n1nn+11(x1+n)n1\displaystyle u(x)\approx\frac{(1+n)^{n+1}}{n^{n-1}}\,\frac{1}{(x-1)^{n+1}},\qquad u(x)\approx\frac{(n-1)^{n-1}}{n^{n+1}}\frac{1}{(x-1+n)^{n-1}}

so we have the inverse functions

(u)[x1n,x1(u)1n+((n1)n1nn+11u)1n1x1,x2(u)1+((1+n)n+1nn11u)1n+1(u\to\infty)\quad\left[\begin{array}[]{@{}l@{\quad}l@{}}x\to 1-n,&x^{\infty}_{\mathfrak{1}}(u)\approx 1-n+\left(\frac{(n-1)^{n-1}}{n^{n+1}}\frac{1}{u}\right)^{\frac{1}{n-1}}\\ x\to 1,&x^{\infty}_{\mathfrak{2}}(u)\approx 1+\left(\frac{(1+n)^{n+1}}{n^{n-1}}\frac{1}{u}\right)^{\frac{1}{n+1}}\end{array}\right. (168)

When u=1u_{*}=1, one cannot find the 2n2n solutions of Eq.(43), but fortunately we are only interested in those solutions which also correspond to the limit t1xt_{1}\to x. In this case, instead of a polynomial equation of degree 2n2n, we must solve Eq.(31) which becomes

2x+n1(n+x)x=0\frac{2x+n-1}{(n+x)x}=0 (169)

with only two solutions: x=x=\infty and x=12(1n)x=\frac{1}{2}(1-n). The behavior of u(x)u(x) near x=x=\infty can be found with the conformal transformation x=1/εx=1/\varepsilon; evaluating u(1/ε)u(1/\varepsilon) around small ε\varepsilon,

u(1/ε)=1+4nε+2n(1+3n)ε2+O(ε3)\displaystyle u(1/\varepsilon)=1+4n\varepsilon+2n(1+3n)\varepsilon^{2}+\mathrm{O}(\varepsilon^{3})

while expanding u(x)u(x) the second limit, when x1n2x\to\frac{1-n}{2}, we get

u(x)=164n3(n21)2(x1n2)3512n(1+n2)5(n21)4(x1n2)5+2048n29(n21)4(x1n2)6+O(x1n2)7\begin{split}u(x)&=1-\tfrac{64n}{3(n^{2}-1)^{2}}\left(x-\tfrac{1-n}{2}\right)^{3}-\tfrac{512n(1+n^{2})}{5(n^{2}-1)^{4}}\left(x-\tfrac{1-n}{2}\right)^{5}\\ &\qquad+\tfrac{2048n^{2}}{9(n^{2}-1)^{4}}\left(x-\tfrac{1-n}{2}\right)^{6}+{\mathrm{O}}\left(x-\tfrac{1-n}{2}\right)^{7}\end{split} (170)

Inverting the two series above, we find

(u1)[x,x11(u)4n1u+3n+12+x1n2,x21(u)1n2+31/3(n21)2/34n1/3(1u)133(n2+1)40n(1u)+(n21)2/3832/3n1/3(1u)43+\displaystyle(u\to 1)\quad\left[\begin{array}[]{@{}l@{\quad}l@{}}x\to\infty,&\quad x^{1}_{\mathfrak{1}}(u)\approx-{4n\over 1-u}+\frac{3n+1}{2}+\cdots\\ \\ x\to\tfrac{1-n}{2},&\quad x^{1}_{\mathfrak{2}}(u)\approx\frac{1-n}{2}\\ &\quad\quad\quad\quad+\tfrac{3^{1/3}(n^{2}-1)^{2/3}}{4n^{1/3}}(1-u)^{\frac{1}{3}}\\ &\quad\quad\quad\quad-\tfrac{3(n^{2}+1)}{40n}(1-u)\\ &\quad\quad\quad\quad+\tfrac{(n^{2}-1)^{2/3}}{8\cdot 3^{2/3}\cdot n^{1/3}}(1-u)^{\frac{4}{3}}+\cdots\end{array}\right. (177)

Note that the multiplicity of the solution x=1n2x=\frac{1-n}{2} is 3, and that of x=x=\infty is 1.

Appendix C OPEs with bare twists and structure constants

In this appendix, we examine the OPE limits of the functions Gσ(x)G_{\sigma}(x) and g(x)g(x). We derive several structure constants, some of which are known in the literature, thus checking our expressions for g(x)g(x) and Gσ(x)G_{\sigma}(x).


We start with the limit u1u\to 1 for G(u,u¯)G(u,\bar{u}). The identity channel is the same as for the Ramond fields discussed in the text, while the second channel gives

Gσ(x21(u))=d4(1u)4/3+d2(1u)2/3+d1(1u)1/3+non-sing.\displaystyle\begin{split}G_{\sigma}(x^{1}_{\mathfrak{2}}(u))&=\frac{d_{4}}{(1-u)^{4/3}}+\frac{d_{2}}{(1-u)^{2/3}}+\frac{d_{1}}{(1-u)^{1/3}}+\text{non-sing.}\end{split} (178)

where, after taking (70) into account,

log|d4|2=(n+1n+23)log(n+1)(n+1n23)log(n1)43logn4log283log3.\displaystyle\begin{split}\log|d_{4}|^{2}&=\left(n+\frac{1}{n}+\frac{2}{3}\right)\log(n+1)-\left(n+\frac{1}{n}-\frac{2}{3}\right)\log(n-1)-\frac{4}{3}\log n\\ &\quad-4\log 2-\frac{8}{3}\log 3.\end{split} (179)

The powers of uu reveal the conformal family of σ3\sigma_{3} with no operator of dimension one among the descendants. Inserting the OPE (72) back into the four-point function, we find that

|d4|2=O2(int)()σ3(1)O2(int)(0)σn()σ3(1)σn(0).\displaystyle|d_{4}|^{2}=\big{\langle}O^{(\mathrm{int})}_{2}(\infty)\sigma_{3}(1)O^{(\mathrm{int})}_{2}(0)\big{\rangle}\,\big{\langle}\sigma_{n}(\infty)\sigma_{3}(1)\sigma_{n}(0)\big{\rangle}. (180)

The structure constant C232O(int)σO(int)=O2(int)()σ3(1)O2(int)(0)C^{O^{(\mathrm{int})}\sigma O^{(\mathrm{int})}}_{232}=\langle O^{(\mathrm{int})}_{2}(\infty)\sigma_{3}(1)O^{(\mathrm{int})}_{2}(0)\rangle carries no nn-dependence, and we can write Cn3nσ=σn()σ3(1)σn(0)C^{\sigma}_{n3n}=\langle\sigma_{n}(\infty)\sigma_{3}(1)\sigma_{n}(0)\rangle as

logCn3nσ=(n+1n+23)log(n+1)(n+1n23)log(n1)43logn+κ,\log C^{\sigma}_{n3n}=\left(n+\frac{1}{n}+\frac{2}{3}\right)\log(n+1)-\left(n+\frac{1}{n}-\frac{2}{3}\right)\log(n-1)-\frac{4}{3}\log n+\kappa, (181)

which agrees with Eq.(6.25) of Ref.Lunin:2000yv , apart from an overall factor of 16\frac{1}{6}. The nn-independent number κ=43log3+13log2\kappa=\tfrac{4}{3}\log 3+\tfrac{1}{3}\log 2 can be obtained by taking n=2n=2 in Eq.(181), and comparing with Eq.(187), below. Hence

logC232O(int)σO(int)=4log3+133log2.\log C^{O^{(\mathrm{int})}\sigma O^{(\mathrm{int})}}_{232}=4\log 3+\tfrac{13}{3}\log 2. (182)

An interesting check of the results above comes from the function (42), whose limit u1u\to 1 now gives the OPE σ2(u)σ2(1)\sigma_{2}(u)\sigma_{2}(1). Counting powers of uu in

g(x11(u))\displaystyle g(x^{1}_{\mathfrak{1}}(u)) =cσ(4n)3/4(1u)3/4+O(1u)1/4\displaystyle=\frac{c_{\sigma}(-4n)^{3/4}}{(1-u)^{3/4}}+\mathrm{O}(1-u)^{1/4} (183)
g(x21(u))\displaystyle g(x^{1}_{\mathfrak{2}}(u)) =b(1u)112+O(1u)7/12\displaystyle=\frac{b}{(1-u)^{1\over 12}}+\mathrm{O}(1-u)^{7/12} (184)

Eq.(183) gives again the identity, thus determining cσ=(4n)3/4c_{\sigma}=(-4n)^{-3/4}. In the channel (184) we find σ3\sigma_{3} with its dimension h3σ=23h^{\sigma}_{3}=\frac{2}{3}, completing the well-known fusion rule [σ2]×[σ2]=[σ1]+[σ3][\sigma_{2}]\times[\sigma_{2}]=[\sigma_{1}]+[\sigma_{3}]. Note the absence of next-to leading singularities in Eqs.(183) and (184); there are no descendants in these OPEs.

The constant bb gives information about σpσqσr=Cpqrσ\langle\sigma_{p}\sigma_{q}\sigma_{r}\rangle=C^{\sigma}_{pqr}:

logCn3nσ+logC232σ\displaystyle\log C^{\sigma}_{n3n}+\log C^{\sigma}_{232} =log|b|2\displaystyle=\log|b|^{2}
=(n+1n+23)log(n+1)(n+1n23)log(n1)\displaystyle=\left(n+\frac{1}{n}+\frac{2}{3}\right)\log(n+1)-\left(n+\frac{1}{n}-\frac{2}{3}\right)\log(n-1)
43logn4log216log3\displaystyle\qquad-\frac{4}{3}\log n-4\log 2-\frac{1}{6}\log 3 (185)

Comparison with Eqs.(180) and (179) reveals the same nn-dependence for both three-point functions — an important cross-check between the two g(x)g(x) and Gσ(x)G_{\sigma}(x) (which were obtained independently).

The more general fusion rule

[σ2]×[σn]=[σn1]+[σn+1][\sigma_{2}]\times[\sigma_{n}]=[\sigma_{n-1}]+[\sigma_{n+1}] (186)

can be derived from

g(x10(u))=u5n25n+28n(n1)(c1+c2u1n1+),g(x20(u))=u5n2+5n28n(n+1)(c1++c2+u1n+1+)\displaystyle g(x^{0}_{\mathfrak{1}}(u))=u^{-\frac{5n^{2}-5n+2}{8n(n-1)}}\big{(}c_{1}^{-}+c_{2}^{-}u^{\frac{1}{n-1}}+\cdots\big{)},\quad g(x^{0}_{\mathfrak{2}}(u))=u^{-\frac{5n^{2}+5n-2}{8n(n+1)}}\big{(}c_{1}^{+}+c_{2}^{+}u^{\frac{1}{n+1}}+\cdots\big{)}

where the coefficients c1±c_{1}^{\pm} are readily computable. One can check from the powers of uu that channels x10(u)x^{0}_{\mathfrak{1}}(u) and x20(u)x^{0}_{\mathfrak{2}}(u) give operators of dimensions hn1σh^{\sigma}_{n-1} and hn+1σh^{\sigma}_{n+1}, that is σn1\sigma_{n-1} and σn+1\sigma_{n+1}, respectively. The coefficients c1±c_{1}^{\pm} give us information about σn()σ2(1)σn1(0)=|c1|\big{\langle}\sigma_{n}(\infty)\sigma_{2}(1)\sigma_{n-1}(0)\big{\rangle}=|c_{1}^{-}| and σn()σ2(1)σn+1(0)=|c1+|\big{\langle}\sigma_{n}(\infty)\sigma_{2}(1)\sigma_{n+1}(0)\big{\rangle}=|c_{1}^{+}|. Explicitly

logCn,2,n1σ\displaystyle\log C^{\sigma}_{n,2,n-1} =2n2n+24nlog(n1)+2n23n+34(n1)logn54log2\displaystyle=-\frac{2n^{2}-n+2}{4n}\log(n-1)+\frac{2n^{2}-3n+3}{4(n-1)}\log n-\frac{5}{4}\log 2 (187)
logCn,2,n+1σ\displaystyle\log C^{\sigma}_{n,2,n+1} =+2n2+n+24nlog(n+1)2n2+3n+34n(n+1)logn54log2\displaystyle=+\frac{2n^{2}+n+2}{4n}\log(n+1)-\frac{2n^{2}+3n+3}{4n(n+1)}\log n-\frac{5}{4}\log 2 (188)

This agrees with the result of Ref.Lunin:2000yv , again, apart from an overall factor of 16\frac{1}{6}. Inserting n=3n=3 in Eq.(187), we find that logC3,2,2σ=53log2\log C^{\sigma}_{3,2,2}=-\frac{5}{3}\log 2, which was used to derived Eq.(182).

References