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

Conformal field theory on the Riemann sphere and its boundary version for SLE

Nam-Gyu Kang School of Mathematics, Korea Institute for Advanced Study,
Seoul, 02455, Republic of Korea
[email protected]
 and  Nikolai G. Makarov Department of Mathematics, California Institute of Technology,
Pasadena, CA 91125, USA
[email protected]
Abstract.

From conformal field theory on the Riemann sphere, we implement its boundary version in a simply-connected domain using the Schottky double construction. We consider the statistical fields generated by background charge modification of the Gaussian free field with Dirichlet boundary condition under the OPE multiplications. We prove that the correlation functions of such fields with symmetric background charges form a collection of martingale-observables for (forward) chordal/radial SLE with force points and spins. We also present the connection between conformal field theory with Neumann boundary condition and the theory of backward SLE.

Key words and phrases:
conformal field theory, Gaussian free field, martingale-observables, SLE
2010 Mathematics Subject Classification:
Primary 60J67, 81T40; Secondary 30C35
Nam-Gyu Kang was partially supported by Samsung Science and Technology Foundation (SSTF-BA1401-51), a KIAS Individual Grant (MG058103) at Korea Institute for Advanced Study, and the National Research Foundation of Korea (NRF-2019R1A5A1028324).
Nikolai G. Makarov was supported by NSF grant no. 1500821.

1. Introduction and results

Since Belavin, Polyakov, and Zamolodchikov [7] introduced an operator algebra formalism to relate some critical models to the representation theory of Virasoro algebra, conformal field theory has been applied in string theory and condensed matter physics. In mathematics, conformal field theory inspired the development of algebraic theories, e.g., the theory of vertex algebras. Also, it has been used to derive several exact results for the conformally invariant critical clusters in two-dimensional lattice models.

During the last two decades, some outstanding predictions from statistical physics have been proved using Schramm-Loewner evolution (SLE). For instance, the remarkable achievements in this area include the work of Lawler-Schramm-Werner on Mandelbrot conjecture (see [23] and references therein), Smirnov’s work on percolation model and Ising model at criticality (see [35, 36]), and the work of Schramm-Sheffield on contour lines of the two-dimensional discrete Gaussian free field (see [31]). Under quite general conditions, it requires just one martingale-observable, which determines the law uniquely to prove the scaling limit convergence of interface curves in lattice models.

In this paper we use the method of conformal field theory to study a certain family of chordal/radial SLE martingale-observables. To be precise, we develop a version of conformal field theory on the Riemann sphere and apply the Schottky double construction to this theory to implement its boundary version in a simply-connected domain DD with marked boundary/interior points. This theory is based on non-random (background charge) modification of the Gaussian free field Φ\Phi in DD with Dirichlet boundary condition. The theory with Neumann boundary condition is presented in Section 9. We state Ward’s equations in terms of Lie derivatives for the fields in the OPE family 𝜷\mathcal{F}_{\boldsymbol{\beta}} generated by the modified Gaussian free field Φ𝜷\Phi_{\boldsymbol{\beta}} with a (symmetric) background charge 𝜷\boldsymbol{\beta} placed at the marked points. Using the BPZ-Cardy equations for the fields in 𝜷\mathcal{F}_{\boldsymbol{\beta}} for a symmetric background charge 𝜷\boldsymbol{\beta} (so that the associated partition function Z𝜷Z_{\boldsymbol{\beta}} is real-valued) with a specific charge at the growth point pp of SLE, we show that the correlation functions of the fields in 𝜷\mathcal{F}_{\boldsymbol{\beta}} form a collection of martingale-observables for chordal/radial SLE[𝜷]\mathrm{SLE}[\boldsymbol{\beta}] associated with the partition function Z𝜷.Z_{\boldsymbol{\beta}}.

The partition function Z𝜷Z_{\boldsymbol{\beta}} we consider is defined in terms of the Coulomb gas correlation function C[𝜷]C[\boldsymbol{\beta}] or the correlation function of the effective one-leg operator. In the physics literature, the partition function in conformal field theory is expressed as the correlation function of the boundary condition changing operator, see ([11, §11.3.1]). The other approaches to define the partition function in the context of SLE (up to a multiplicative constant depending possibly on DD) can be found in [14, 22]. After introducing the partition functions of free field and the partition functions of SLE, Dubédat showed that they coincide and use these identities to prove local/global couplings of SLEs and free fields in [14]. The SLE partition function can be viewed as the total mass of natural measure on SLE curves and used to describe the restriction property, e.g., see [22].

It is a well-known statement in the physics literature (e.g., see [9, 12]) that under the insertion of the normalized one-leg operator Λp/(𝐄Λp)\Lambda_{p}/(\mathbf{E}\,\Lambda_{p}) rooted at qDDq\in D\cup\partial D all correlation functions of fields in the OPE family 𝜷ˇ\mathcal{F}_{\check{\boldsymbol{\beta}\,}} are martingale-observables for SLE(κ)\mathrm{SLE}(\kappa) in DD from pDp\in\partial D to q.q. (Here 𝐄Λp\mathbf{E}\,\Lambda_{p} is the correlation function of Λp\Lambda_{p} and a specific background charge 𝜷ˇ\check{\boldsymbol{\beta}\,} is placed at q.q.) We construct the one-leg “operator” Λp\Lambda_{p} rooted at qq as one of the OPE exponentials of the linear combination of bosonic fields with proper charges at p,qp,q so that it is a primary field with the desired conformal dimensions at pp and qq producing a level two degenerate vector at p.p. Considering general non-random pre-pre-Schwarzian modifications of the Gaussian free field, we extend this result to the chordal/radial SLE with forces and spins.

In constructing such OPE exponentials, it is essential to require the so-called neutrality conditions on the charges. We clarify and explain two neutrality conditions: one condition for the linear combination of bosonic fields to be a well-defined (Fock space) field; the other condition for the (Coulomb gas) correlation function to be conformally invariant. To reconcile these two neutrality conditions, we place the “background charge(s)” 𝜷\boldsymbol{\beta} at the marked point(s). We also derive the neutrality condition on the background charges using an extension of the Gauss-Bonnet theorem to the flat metric with conical singularities at the marked points. It appears that our approach can be extended to much more general settings like general Riemann surfaces with marked points (e.g., see [8] for this approach in a doubly-connected domain) and various patterns of insertion (e.g., NN-leg operators with the method of screening, see [1]). However, we do not claim that what we develop here is the only relevant conformal field theory to SLE theory. For example, as another relevant theory, a twisted conformal field theory can be related to radial SLE. Its bosonic observable with central charge c=1c=1 corresponds to a discrete observable of the harmonic explorer with 2\mathbb{Z}_{2}-symmetry.

1.1. A spherical conformal field theory and its boundary version

The fields we treat in this paper are statistical fields constructed from the Gaussian free field Φ\Phi with Dirichlet boundary condition via differentiation and Wick’s formula. In the last section we develop a version of conformal field theory constructed from the Gaussian free field NN with Neumann boundary condition. The Gaussian free field Φ\Phi in a simply-connected domain DD can be viewed as a Fock space field and its 2-point correlation is given by

𝐄Φ(z1)Φ(z2)=2G(z1,z2),\mathbf{E}\,\Phi(z_{1})\Phi(z_{2})=2G(z_{1},z_{2}),

where GG is the Dirichlet Green’s function for D.D. A Fock space field is a linear combination of basic fields, which are, by definition, formal expressions written as Wick’s product (\odot-product) of derivatives of Φ\Phi and Wick’s exponentials eαΦ(α)\mathrm{e}^{\odot\alpha\Phi}\,(\alpha\in\mathbb{C}) of Φ.\Phi. If X1,,XnX_{1},\cdots,X_{n} are Fock space fields and z1,,znz_{1},\cdots,z_{n} are distinct points in D,D, then a correlation function 𝐄[X1(z1)Xn(zn)]\mathbf{E}[X_{1}(z_{1})\cdots X_{n}(z_{n})] can be defined by Wick’s calculus. See Section 3 or [18, Lecture 1] for more details. (A more traditional notation for the correlation function is X1(z1)Xn(zn).\langle X_{1}(z_{1})\cdots X_{n}(z_{n})\rangle.) The points z1,,znz_{1},\cdots,z_{n} are called the nodes of the (tensor) product 𝒳=X1(z1)Xn(zn)\mathcal{X}=X_{1}(z_{1})\cdots X_{n}(z_{n}) of Xj(zj)X_{j}(z_{j})’s and we write S𝒳S_{\mathcal{X}} for {z1,,zn}.\{z_{1},\cdots,z_{n}\}. We extend the collection of Fock space fields by adding finite linear combinations of formal bosonic fields

Φ[𝝉+,𝝉]:=τj+Φ+(zj)τjΦ(zj)\Phi[\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}]:=\sum\tau_{j}^{+}\Phi^{+}(z_{j})-\tau_{j}^{-}\Phi^{-}(z_{j})

with the neutrality condition

(NC0\mathrm{NC}_{0}) j(τj++τj)=0\sum_{j}(\tau_{j}^{+}+\tau_{j}^{-})=0

and its Wick’s exponential eiΦ[𝝉+,𝝉].\mathrm{e}^{\odot i\Phi[\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}]}. Here 𝝉+\boldsymbol{\tau}^{+} and 𝝉\boldsymbol{\tau}^{-} are divisors in DD which takes the value 0 at all points except for finitely many points:

𝝉±=τj±zj.\boldsymbol{\tau}^{\pm}=\sum\tau_{j}^{\pm}\cdot z_{j}.

The centered Gaussian formal fields Φ+,Φ\Phi^{+},\Phi^{-} have formal correlations

𝐄[Φ+(z1)Φ+(z2)]=log1z1z2,𝐄[Φ+(z1)Φ(z2)]=log(1z1z¯2)\mathbf{E}[\Phi^{+}(z_{1})\Phi^{+}(z_{2})]=\log\frac{1}{z_{1}-z_{2}},\qquad\mathbf{E}[\Phi^{+}(z_{1})\Phi^{-}(z_{2})]=\log(1-z_{1}\bar{z}_{2})

in 𝔻\mathbb{D} and the relations: Φ=Φ+¯,Φ=Φ++Φ.\Phi^{-}=\overline{\Phi^{+}},\Phi=\Phi^{+}+\Phi^{-}.

• A spherical conformal field theory.  We consider a compact Riemann surface SS of genus-zero with (finitely many) marked points qk.q_{k}. We introduce the formal fields Ψ±\Psi^{\pm} (Ψ=Ψ+¯\Psi^{-}=\overline{\Psi^{+}}) on SS with formal correlations

𝐄[Ψ+(z1)Ψ+(z2)]=log1z1z2,𝐄[Ψ+(z1)Ψ(z2)]=0\mathbf{E}[\Psi^{+}(z_{1})\Psi^{+}(z_{2})]=\log\frac{1}{z_{1}-z_{2}},\qquad\mathbf{E}[\Psi^{+}(z_{1})\Psi^{-}(z_{2})]=0

on ^.\widehat{\mathbb{C}}. (A bi-variant field Ψ±(z,z0)=Ψ±(z)Ψ±(z0)\Psi^{\pm}(z,z_{0})=\Psi^{\pm}(z)-\Psi^{\pm}(z_{0}) can be defined as a multivalued correlation functional valued field.) We then modify Ψ+\Psi^{+} (and Ψ\Psi^{-} in a similar way) by adding a non-random pre-pre-Schwarzian form (PPS form) ψ+\psi^{+} of order (ib,0)(ib,0) with logarithmic singularities at qkq_{k}’s. Here bb is the modification parameter. Such ψ+\psi^{+} resolves a background charge 𝜷=i¯ψ+/π=βkδk.\boldsymbol{\beta}=i\partial{\bar{\partial}}\psi^{+}/\pi=\sum\beta_{k}\delta_{k}. We write ψ𝜷+=ψ+,\psi^{+}_{\boldsymbol{\beta}}=\psi^{+}, Ψ𝜷+=Ψ++ψ𝜷+.\Psi^{+}_{\boldsymbol{\beta}}=\Psi^{+}+\psi^{+}_{\boldsymbol{\beta}}. A version of the Gauss-Bonnet theorem leads to the neutrality condition (NCb):(\mathrm{NC}_{b}):

kβk=2b.\sum_{k}\beta_{k}=2b.

For 𝝉=τjzj\boldsymbol{\tau}=\sum\tau_{j}\cdot z_{j} satisfying (NC0)(\mathrm{NC}_{0}) we define the OPE exponential 𝒪𝜷[𝝉]\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}] (whose meaning is explained in Subsection 3.4) of Ψ+[𝝉]\Psi^{+}[\boldsymbol{\tau}] by

𝒪𝜷[𝝉]:=C(b)[𝝉+𝜷]C(b)[𝜷]eiΨ+[𝝉],Ψ+[𝝉]:=τj+Ψ+(zj),\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}]:=\frac{C_{(b)}[\boldsymbol{\tau}+\boldsymbol{\beta}]}{C_{(b)}[\boldsymbol{\beta}]}\,\mathrm{e}^{\odot i\Psi^{+}[\boldsymbol{\tau}]},\quad\Psi^{+}[\boldsymbol{\tau}]:=\sum\tau_{j}^{+}\Psi^{+}(z_{j}),

where the (formal) Coulomb gas correlation function C(b)[𝝈]C_{(b)}[\boldsymbol{\sigma}] is a 𝝀\boldsymbol{\lambda}-differential (𝝀=λjzj,λj=σj2/2σjb)(\boldsymbol{\lambda}=\sum\lambda_{j}\cdot z_{j},\lambda_{j}=\sigma_{j}^{2}/2-\sigma_{j}b) such that its evaluation on ^={}\widehat{\mathbb{C}}=\mathbb{C}\cup\{\infty\} is given by

(C(b)[𝝈]id^)=j<k(zjzk)σjσk,(C_{(b)}[\boldsymbol{\sigma}]\,\|\,\mathrm{id}_{\widehat{\mathbb{C}}})=\prod_{j<k}(z_{j}-z_{k})^{\sigma_{j}\sigma_{k}},

where the product is taken over finite zjz_{j}’s and zkz_{k}’s. One can view C(b)[𝝈]C_{(b)}[\boldsymbol{\sigma}] as the formal correlations

C(b)[𝝈]=𝐄eiσ1Ψ+(z1)eiσnΨ+(zn)C_{(b)}[\boldsymbol{\sigma}]=\mathbf{E}\,\mathrm{e}^{\odot i\sigma_{1}\Psi^{+}(z_{1})}\cdots\mathrm{e}^{\odot i\sigma_{n}\Psi^{+}(z_{n})}

by Wick’s rule of centered jointly (generalized) Gaussians:

eα1ξ1eαnξn=ej<kαjαk𝐄[ξjξk]ejαjξj.\mathrm{e}^{\odot\alpha_{1}\xi_{1}}\cdots\mathrm{e}^{\odot\alpha_{n}\xi_{n}}=\mathrm{e}^{\sum_{j<k}\alpha_{j}\alpha_{k}\mathbf{E}[\xi_{j}\xi_{k}]}\mathrm{e}^{\odot\sum_{j}\alpha_{j}\xi_{j}}.

• Schottky double construction.  We consider a simply-connected domain DD with marked boundary/interior points qkD¯q_{k}\in\mkern 1.5mu\overline{\mkern-1.5muD\mkern-1.5mu}\mkern 1.5mu and the Schottky double SS of DD with the canonical involution zz.z\mapsto z^{*}. For a divisor 𝝈=σjzj,\boldsymbol{\sigma}=\sum\sigma_{j}\cdot z_{j}, let us denote 𝝈=σjzj,𝝈¯=σ¯jzj.\boldsymbol{\sigma}_{*}=\sum\sigma_{j}\cdot z_{j}^{*},\mkern 1.5mu\overline{\mkern-1.5mu\boldsymbol{\sigma}\mkern-1.5mu}\mkern 1.5mu=\sum\mkern 1.5mu\overline{\mkern-1.5mu\sigma\mkern-1.5mu}\mkern 1.5mu_{j}\cdot z_{j}. Suppose that a (double) background charge (𝜷+,𝜷)(\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}) (𝜷±=βk±qk)(\boldsymbol{\beta}^{\pm}=\sum\beta_{k}^{\pm}\cdot q_{k}) satisfies the neutrality condition (NCb):(\mathrm{NC}_{b}):

(βk++βk)=2b.\sum(\beta_{k}^{+}+\beta_{k}^{-})=2b.

For this pair (𝜷+,𝜷)(\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}) and a double divisor (𝝉+,𝝉)(\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}) satisfying the neutrality condition (NC0),(\mathrm{NC}_{0}), we define the modification Φ𝜷\Phi_{\boldsymbol{\beta}} of Φ\Phi (𝜷=𝜷++𝜷,𝝉=𝝉++𝝉)(\boldsymbol{\beta}=\boldsymbol{\beta}^{+}+\boldsymbol{\beta}^{-}_{*},\boldsymbol{\tau}=\boldsymbol{\tau}^{+}+\boldsymbol{\tau}^{-}_{*}) by

Φ𝜷[𝝉]=12Ψ2𝜷+,2𝜷[𝝉,𝝉],\Phi_{\boldsymbol{\beta}}[\boldsymbol{\tau}]=\frac{1}{\sqrt{2}}\Psi_{\sqrt{2}\,\boldsymbol{\beta}^{+},\sqrt{2}\,\boldsymbol{\beta}_{*}^{-}}[\boldsymbol{\tau},\boldsymbol{\tau}_{*}],

where Ψ𝜷+,𝜷[𝝉+,𝝉]:=Ψ𝜷++[𝝉+]+Ψ𝜷[𝝉]\Psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}[\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}]:=\Psi_{\boldsymbol{\beta}^{+}}^{+}[\boldsymbol{\tau}^{+}]+\Psi_{\boldsymbol{\beta}^{-}}^{-}[\boldsymbol{\tau}^{-}] and Ψ𝜷[𝝉]=Ψ𝜷¯+[𝝉¯]¯.\Psi_{\boldsymbol{\beta}^{-}}^{-}[\boldsymbol{\tau}^{-}]=\overline{\Psi^{+}_{\mkern 1.5mu\overline{\mkern-1.5mu\boldsymbol{\beta}^{-}\mkern-1.5mu}\mkern 1.5mu}[\overline{\boldsymbol{\tau}^{-}}]}. We also define the multi-vertex field V𝜷[𝜷+𝝉]V_{\boldsymbol{\beta}}[\boldsymbol{\beta}+\boldsymbol{\tau}] and the OPE exponential 𝒪𝜷[𝝉]\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}] by

V𝜷[𝜷+𝝉]=C(b)[𝜷+𝝉]eiΦ[𝝉],𝒪𝜷[𝝉]=C(b)[𝝉+𝜷]C(b)[𝜷]eiΦ[𝝉].V_{\boldsymbol{\beta}}[\boldsymbol{\beta}+\boldsymbol{\tau}]=C_{(b)}[\boldsymbol{\beta}+\boldsymbol{\tau}]\,\mathrm{e}^{\odot i\Phi[\boldsymbol{\tau}]},\qquad\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}]=\frac{C_{(b)}[\boldsymbol{\tau}+\boldsymbol{\beta}]}{C_{(b)}[\boldsymbol{\beta}]}\,\mathrm{e}^{\odot i\Phi[\boldsymbol{\tau}]}.

In [18] we study the standard chordal theory with 𝜷+=2bq,\boldsymbol{\beta}^{+}=2b\cdot q, 𝜷=𝟎\boldsymbol{\beta}^{-}=\boldsymbol{0} (qDq\in\partial D). In this case, the modification Φ𝜷\Phi_{\boldsymbol{\beta}} reads as

Φ𝜷=Φ+φ𝜷,φ𝜷=2bargw,\Phi_{\boldsymbol{\beta}}=\Phi+\varphi_{\boldsymbol{\beta}},\qquad\varphi_{\boldsymbol{\beta}}=-2b\,\arg w^{\prime},

where w:(D,q)(,)w:(D,q)\to(\mathbb{H},\infty) is a conformal map from DD onto the upper half-plane .\mathbb{H}.

For the standard radial theory (𝜷+=𝜷=bq,\boldsymbol{\beta}^{+}=\boldsymbol{\beta}^{-}=b\cdot q, qDq\in D) without further marked points or spins, the modification Φ𝜷\Phi_{\boldsymbol{\beta}} reads as

Φ𝜷=Φ+φ𝜷,φ𝜷=2bargww,\Phi_{\boldsymbol{\beta}}=\Phi+\varphi_{\boldsymbol{\beta}},\qquad\varphi_{\boldsymbol{\beta}}=-2b\,\arg\frac{w^{\prime}}{w},

where w:(D,q)(𝔻,0)w:(D,q)\to(\mathbb{D},0) is a conformal map from DD onto the unit disc 𝔻.\mathbb{D}. The corresponding conformal field theory is Aut(D,q)\mathrm{Aut}(D,q)-invariant. The non-random harmonic function φ𝜷\varphi_{\boldsymbol{\beta}} on a punctured domain D:=D{q}D^{*}:=D\setminus\{q\} is multivalued.

The bosonic fields Φ𝜷[𝝉]\Phi_{\boldsymbol{\beta}}[\boldsymbol{\tau}] with the neutrality condition generate the OPE family 𝜷,\mathcal{F}_{\boldsymbol{\beta}}, the algebra (over \mathbb{C}) spanned by 11 and the derivatives of Φ𝜷,\Phi_{\boldsymbol{\beta}}, 𝒪𝜷[𝝉]\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}] (supp𝜷supp𝝉=\mathrm{supp}\,\boldsymbol{\beta}\cap\mathrm{supp}\,\boldsymbol{\tau}=\emptyset) under the OPE multiplication .*. (If a single-valued field XX is holomorphic, then the OPE product XYX*Y of two fields XX and YY is the zeroth coefficient of the regular part in the operator product expansion X(ζ)Y(z)X(\zeta)Y(z) as ζz.\zeta\to z.) For example, 𝜷\mathcal{F}_{\boldsymbol{\beta}} contains 1,J𝜷:=Φ𝜷,J𝜷(Φ𝜷Φ𝜷),1,J_{\boldsymbol{\beta}}:=\partial\Phi_{\boldsymbol{\beta}},\partial J_{\boldsymbol{\beta}}*(\Phi_{\boldsymbol{\beta}}*\Phi_{\boldsymbol{\beta}}), etc., and the Virasoro field T𝜷,T_{\boldsymbol{\beta}},

T𝜷=12J𝜷J𝜷+ibJ𝜷.T_{\boldsymbol{\beta}}=-\frac{1}{2}J_{\boldsymbol{\beta}}*J_{\boldsymbol{\beta}}+ib\partial J_{\boldsymbol{\beta}}.

In Section 6 we extend the OPE family by adding the OPE functionals at the punctures where the background charges are placed. For example, for 𝜷ˇ=bq+bq,\check{\boldsymbol{\beta}\,}=b\cdot q+b\cdot q^{*}, (qD)(q\in D)

(J𝜷ˇ)q=J(q)+ib2Nw(q),Nw=w′′w(J_{\check{\boldsymbol{\beta}\,}})_{q}=J(q)+\frac{ib}{2}N_{w}(q),\qquad N_{w}=\frac{w^{\prime\prime}}{w^{\prime}}

and 𝒪𝜷[𝝉]\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}] (supp𝜷supp𝝉\mathrm{supp}\,\boldsymbol{\beta}\cap\mathrm{supp}\,\boldsymbol{\tau}\neq\emptyset) belong to the extended OPE family.

• Insertion formulas.  We explain how the insertion of Wick’s exponential V[𝝉]:=eiΦ[𝝉]V^{\odot}[\boldsymbol{\tau}]:=\mathrm{e}^{\odot i\Phi[\boldsymbol{\tau}]} of the Gaussian free field gives rise to the change of background charge modifications. A particular insertion procedure plays an essential role in establishing the connection between chordal/radial SLE theory with force points (and spins) and conformal field theory in a simply-connected domain with certain symmetric background charges.

Theorem 1.1.

Given two background charges 𝛃ˇ,𝛃\check{\boldsymbol{\beta}\,},\boldsymbol{\beta} with the neutrality condition (NCb),(\mathrm{NC}_{b}), the image of 𝛃ˇ\mathcal{F}_{\check{\boldsymbol{\beta}\,}} under the insertion of V[𝛃𝛃ˇ]V^{\odot}[\boldsymbol{\beta}-\check{\boldsymbol{\beta}\,}] is 𝛃.\mathcal{F}_{\boldsymbol{\beta}}.

• Ward’s equations and BPZ-Cardy equations.  We mostly concern ourselves with a symmetric background charge 𝜷=𝜷¯.\boldsymbol{\beta}=\mkern 1.5mu\overline{\mkern-1.5mu\boldsymbol{\beta}_{*}\mkern-1.5mu}\mkern 1.5mu. One of our main goals is to present the precise relation between conformal field theory 𝜷\mathcal{F}_{\boldsymbol{\beta}} and SLE[𝜷]\mathrm{SLE}[\boldsymbol{\beta}] for such a symmetric background charge 𝜷,\boldsymbol{\beta}, see Theorem 1.5 below.

Let A𝜷:=T𝜷𝐄T𝜷.A_{\boldsymbol{\beta}}:=T_{\boldsymbol{\beta}}-\mathbf{E}\,T_{\boldsymbol{\beta}}. For a meromorphic vector field in D¯,\bar{D}, we define

W𝜷(v)=2ReW𝜷+(v),W𝜷+(v)=12πiDvA𝜷1πD(¯v)A𝜷,W_{\boldsymbol{\beta}}(v)=2\,\mathrm{Re}\,W_{\boldsymbol{\beta}}^{+}(v),\qquad W_{\boldsymbol{\beta}}^{+}(v)=\frac{1}{2\pi i}\int_{\partial D}vA_{\boldsymbol{\beta}}-\frac{1}{\pi}\int_{D}({\bar{\partial}}v)A_{\boldsymbol{\beta}},

where ¯v{\bar{\partial}}v in the last integral should be understood in the sense of distributions and the first integral should be taken in the sense of the Cauchy principal value if necessary. The following theorem enables us to represent the action of the Lie derivatives operators v\mathcal{L}_{v} by inserting the Ward functionals into correlation functions of fields in the extended OPE family.

Theorem 1.2 (Ward’s identities).

Let X𝛃X_{\boldsymbol{\beta}} be a tensor product of fields in the extended OPE family 𝛃.\mathcal{F}_{\boldsymbol{\beta}}. If the nodes of X𝛃X_{\boldsymbol{\beta}} do not intersect with the poles of v,v, then we have

𝐄vX𝜷=𝐄W𝜷(v)X𝜷.\mathbf{E}\,\mathcal{L}_{v}X_{\boldsymbol{\beta}}=\mathbf{E}\,W_{\boldsymbol{\beta}}(v)X_{\boldsymbol{\beta}}.

The proof of Theorem 1.2 is based on the following residue form of Ward identity

(1.1) v+(z)X𝜷=12πi(z)vA𝜷X𝜷\mathcal{L}_{v}^{+}(z)X_{\boldsymbol{\beta}}=\frac{1}{2\pi i}\oint_{(z)}vA_{\boldsymbol{\beta}}X_{\boldsymbol{\beta}}

for a field X𝜷X_{\boldsymbol{\beta}} in the OPE family with a proper meaning of equality. Here v+\mathcal{L}_{v}^{+} is the \mathbb{C}-linear part of the Lie derivative v.\mathcal{L}_{v}. We emphasize that (1.1) has a different nature at zz in supp𝜷+supp𝜷\mathrm{supp}\,\boldsymbol{\beta}^{+}\cup\mathrm{supp}\,\boldsymbol{\beta}^{-} and at each node zz of X𝜷.X_{\boldsymbol{\beta}}. As Fock space functionals, (1.1) holds at each node zz of X𝜷,X_{\boldsymbol{\beta}}, see Lemma 7.1. On the other hand, (1.1) holds within correlations at zz in supp𝜷+supp𝜷,\mathrm{supp}\,\boldsymbol{\beta}^{+}\cup\mathrm{supp}\,\boldsymbol{\beta}^{-}, see Lemma 7.2.

Ward’s equations (see Theorem 1.3 below) describe the insertion of the Virasoro field within correlations of fields in 𝜷\mathcal{F}_{\boldsymbol{\beta}} in terms of the Lie derivative operators and the puncture operators 𝒫𝜷:=C(b)[𝜷].\mathcal{P}_{\boldsymbol{\beta}}:=C_{(b)}[\boldsymbol{\beta}]. Let kzk_{z} (resp. vzv_{z}) be the chordal (resp. radial) Loewner vector field with pole at z:z:

kz(ζ)=1zζ,(resp. vz(ζ)=ζz+ζzζ)k_{z}(\zeta)=\frac{1}{z-\zeta},\quad\Big{(}\textrm{resp. }v_{z}(\zeta)=\zeta\frac{z+\zeta}{z-\zeta}\Big{)}

in the identity chart of \mathbb{H} (resp. in the identity chart of 𝔻\mathbb{D}).

Theorem 1.3 (Ward’s equations).

Given a symmetric background charge 𝛃\boldsymbol{\beta} with the neutrality condition (NCb),(\mathrm{NC}_{b}), let X1,,Xn𝛃X_{1},\cdots,X_{n}\in\mathcal{F}_{\boldsymbol{\beta}} and let XX be the tensor product of XjX_{j}’s. Then

𝐄T𝜷(z)X=𝐄𝒫𝜷1kz+𝒫𝜷X+𝐄𝒫𝜷1kz¯𝒫𝜷X,\mathbf{E}\,T_{\boldsymbol{\beta}}(z)\,X=\mathbf{E}\,\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{+}_{k_{z}}\mathcal{P}_{\boldsymbol{\beta}}X+\mathbf{E}\,\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{-}_{k_{\bar{z}}}\mathcal{P}_{\boldsymbol{\beta}}X,

where all fields are evaluated in the identity chart of \mathbb{H} and

2z2𝐄T𝜷(z)X=𝐄𝒫𝜷1vz+𝒫𝜷X+𝐄𝒫𝜷1vz𝒫𝜷X,2z^{2}\mathbf{E}\,T_{\boldsymbol{\beta}}(z)\,X=\mathbf{E}\,\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{+}_{v_{z}}\mathcal{P}_{\boldsymbol{\beta}}X+\mathbf{E}\,\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{-}_{v_{z^{*}}}\mathcal{P}_{\boldsymbol{\beta}}X,

where all fields are evaluated in the identity chart of 𝔻.\mathbb{D}.

Let 𝜷\boldsymbol{\beta} be a symmetric background charge with the neutrality condition (NCb)(\mathrm{NC}_{b}) and a specific charge aa at a marked boundary point pD.p\in\partial D. We consider a symmetric divisor 𝝉=ap+\boldsymbol{\tau}=a\cdot p+\cdots satisfying the neutrality condition (NC0)(\mathrm{NC}_{0}) and let 𝜷ˇ=𝜷𝝉.\check{\boldsymbol{\beta}\,}=\boldsymbol{\beta}-\boldsymbol{\tau}. By Theorem 1.1, 𝜷\mathcal{F}_{\boldsymbol{\beta}} is the image of 𝜷ˇ\mathcal{F}_{\check{\boldsymbol{\beta}\,}} under the insertion of

Λp𝐄Λp,\frac{\Lambda_{p}}{\mathbf{E}\,\Lambda_{p}},

where Λp:=𝒪𝜷ˇ[𝝉]\Lambda_{p}:=\mathcal{O}_{\check{\boldsymbol{\beta}\,}}[\boldsymbol{\tau}] is the one-leg operator and 𝐄Λp=C(b)[𝜷]/C(b)[𝜷ˇ]\mathbf{E}\,\Lambda_{p}=C_{(b)}[\boldsymbol{\beta}]/C_{(b)}[\check{\boldsymbol{\beta}}] is its correlation. Let 𝒳\mathcal{X} be a string in 𝜷ˇ.\mathcal{F}_{\check{\boldsymbol{\beta}\,}}. We assume that ζ\zeta is not a node of 𝒳.\mathcal{X}. Denote

𝐄^ζ𝒳:=𝐄Λζ𝒳𝐄Λζ.\widehat{\mathbf{E}}_{\zeta}\,\mathcal{X}:=\frac{\mathbf{E}\,\Lambda_{\zeta}\mathcal{X}}{\mathbf{E}\,\Lambda_{\zeta}}.

We define the partition function associated with a symmetric background charge 𝜷\boldsymbol{\beta} by

Z𝜷:=|C(b)[𝜷]|Z_{\boldsymbol{\beta}}:=\big{|}C_{(b)}[\boldsymbol{\beta}]\big{|}

and let Zζ=Z𝜷ζ,𝜷ζ=𝜷+aζap.Z_{\zeta}=Z_{\boldsymbol{\beta}_{\zeta}},\boldsymbol{\beta}_{\zeta}=\boldsymbol{\beta}+a\cdot\zeta-a\cdot p. The following version of BPZ-Cardy equations play a crucial role in understanding precise relation between conformal field theory and SLE.

Theorem 1.4.

Given a symmetric background charge 𝛃\boldsymbol{\beta} with 𝛃(p)=a\boldsymbol{\beta}(p)=a and the neutrality condition (NCb),(\mathrm{NC}_{b}), let X1,,Xn𝛃X_{1},\cdots,X_{n}\in\mathcal{F}_{\boldsymbol{\beta}} and let XX be the tensor product of XjX_{j}’s. Suppose that the parameters aa and bb satisfy

2a(a+b)=1.2a(a+b)=1.

If ξ,\xi\in\mathbb{R}, then in the identity chart of ,\mathbb{H}, we have

12a2(ξ2+2(ξlogZξ)ξ)𝐄^ξX=ˇkξ𝐄^ξX,\frac{1}{2a^{2}}\Big{(}\partial_{\xi}^{2}+2\big{(}\partial_{\xi}\log Z_{\xi}\big{)}\partial_{\xi}\Big{)}\widehat{\mathbf{E}}_{\xi}\,X=\check{\mathcal{L}}_{k_{\xi}}\widehat{\mathbf{E}}_{\xi}\,X,

where ξ=+¯\partial_{\xi}=\partial+{\bar{\partial}} is the operator of differentiation with respect to the real variable ξ\xi and ˇkξ\check{\mathcal{L}}_{k_{\xi}} is taken over the finite notes of XX and supp𝛃ξ{ξ}.\mathrm{supp}\,\boldsymbol{\beta}_{\xi}\setminus\{\xi\}. In the (𝔻,0)(\mathbb{D},0)-uniformization we have

2a2(12θ2+(θlogZζ)θ)𝐄^ζX=ˇvζ𝐄^ζX,(ζ=eiθ,θ),-\frac{2}{a^{2}}\Big{(}\frac{1}{2}\partial_{\theta}^{2}+\big{(}\partial_{\theta}\log Z_{\zeta}\big{)}\partial_{\theta}\Big{)}\widehat{\mathbf{E}}_{\zeta}X=\check{\mathcal{L}}_{v_{\zeta}}\widehat{\mathbf{E}}_{\zeta}X,\qquad(\zeta=\mathrm{e}^{i\theta},\theta\in\mathbb{R}),

where the Lie derivative ˇvζ\check{\mathcal{L}}_{v_{\zeta}} is taken over the finite notes of XX and supp𝛃ζ{ζ}.\mathrm{supp}\,\boldsymbol{\beta}_{\zeta}\setminus\{\zeta\}.

1.2. SLE and martingale-observables

Since Schramm introduced SLE in [30] as the only possible candidates for the scaling limits of interface curves in critical 2-D lattice models, SLE has been used with remarkable success to prove some important conjectures in statistical physics. For example, see the work of Lawler-Schramm-Werner ([23, 25]), Schramm-Sheffield ([31, 32]), and Smirnov ([35, 36]).

• Standard chordal/radial SLEs.  For a simply-connected domain (D,p,q)(D,p,q) with a marked boundary point pp and a marked interior point q,q, radial Schramm-Loewner evolution (SLE) in (D,p,q)(D,p,q) with a positive parameter κ\kappa is the conformally invariant law on random curves from pp to qq satisfying the so-called “domain Markov property” (see (1.3) below). Let us review the basic definitions of radial SLE(κ)\mathrm{SLE}(\kappa) in technical terms. For each zD,z\in D, let gt(z)g_{t}(z) be the solution (which exists up to a time τz(0,]\tau_{z}\in(0,\infty]) of the equation

(1.2) tgt(z)=gt(z)ζt+gt(z)ζtgt(z),\partial_{t}g_{t}(z)=g_{t}(z)\frac{\zeta_{t}+g_{t}(z)}{\zeta_{t}-g_{t}(z)},

where g0:(D,p,q)(𝔻,1,0)g_{0}:(D,p,q)\to(\mathbb{D},1,0) is a given conformal map, ζt=eiθt,θt=κBt,\zeta_{t}=\mathrm{e}^{i\theta_{t}},\theta_{t}=\sqrt{\kappa}B_{t}, and BtB_{t} is a standard Brownian motion with B0=0.B_{0}=0. Then for all t,t,

wt:(Dt,γt,q)(𝔻,1,0),wt(z):=gt(z)/ζt=gt(z)eiκBtw_{t}:(D_{t},\gamma_{t},q)\to(\mathbb{D},1,0),\qquad w_{t}(z):=g_{t}(z)/\zeta_{t}=g_{t}(z)\mathrm{e}^{-i\sqrt{\kappa}B_{t}}

is a well-defined conformal map from Dt:={zD:τz>t}D_{t}:=\{z\in D:\tau_{z}>t\} onto the unit disc 𝔻.\mathbb{D}. It is known that the SLE stopping time τz\tau_{z} (defined to be the first time when the solution of Loewner equation (1.2) does not exist) satisfies limtτzwt(z)=1.\lim_{t\uparrow\tau_{z}}w_{t}(z)=1. The radial SLE curve γ\gamma is defined by the equation

γtγ(t):=limz1wt1(z)\gamma_{t}\equiv\gamma(t):=\lim_{z\to 1}w_{t}^{-1}(z)

and satisfies the “domain Markov property,”

(1.3) law(γ[t,)|γ[0,t])=lawγDt,γt,q[0,).\mathrm{law}\,\left(\gamma[t,\infty)\,|\,\gamma[0,t]\right)\,=\,\mathrm{law}\,\gamma_{D_{t},\gamma_{t},q}[0,\infty).

The sets Kt:={z𝔻¯:τzt}K_{t}:=\{z\in\overline{\mathbb{D}}:\tau_{z}\leq t\} are called the hulls of the SLE.

For a simply-connected domain (D,p,q)(D,p,q) with two marked boundary points p,q,p,q, the chordal SLE map gtg_{t} with a positive parameter κ\kappa satisfies the equation

(1.4) tgt(z)=2gt(z)ξt,\partial_{t}g_{t}(z)=\frac{2}{g_{t}(z)-\xi_{t}},

where g0:(D,p,q)(,0,)g_{0}:(D,p,q)\to(\mathbb{H},0,\infty) is a given conformal map and ξt=κBt.\xi_{t}=\sqrt{\kappa}B_{t}. Then for all t,t,

wt(z):=gt(z)ξtw_{t}(z):=g_{t}(z)-\xi_{t}

is a well-defined conformal map from the domain Dt:={zD:τz>t}D_{t}:=\{z\in D:\tau_{z}>t\} onto ,\mathbb{H}, where the SLE stopping time τz\tau_{z} satisfies limtτzwt(z)=0.\lim_{t\uparrow\tau_{z}}w_{t}(z)=0. The chordal SLE curve γ\gamma is defined by the equation

γtγ(t):=limz0wt1(z).\gamma_{t}\equiv\gamma(t):=\lim_{z\to 0}w_{t}^{-1}(z).

• Martingale-observables.  Many results in the SLE theory and its applications depend on the explicit form of certain martingale-observables. A non-random conformal field ff is an assignment of a (smooth) function (fϕ):ϕU(f\,\|\,\phi):~{}\phi U\to\mathbb{C} to each local chart ϕ:UϕU.\phi:U\to\phi U. For a non-random conformal field MM of nn variables either in \mathbb{H} (in the chordal case) or in 𝔻\mathbb{D} (in the radial case), let us define MM in any simply-connected domain (D,p,q)(D,p,q) with marked points pD,qDDp\in\partial D,q\in\partial D\cup D by

(MD,p,qid)=(Mw1).(M_{D,p,q}\,\|\,\mathrm{id})=(M\,\|\,w^{-1}).

Here w:(D,p,q)(,0,)w:(D,p,q)\to(\mathbb{H},0,\infty) is a conformal map in the chordal case (qDq\in\partial D) and w:(D,p,q)(𝔻,1,0)w:(D,p,q)\to(\mathbb{D},1,0) is a conformal map in the radial case (qDq\in D). We say that MM is a martingale-observable for chordal/radial SLE(κ)\mathrm{SLE}(\kappa) if for any z1,,znD,z_{1},\cdots,z_{n}\in D, the process

Mt(z1,,zn)=MDt,γt,q(z1,,zn)M_{t}(z_{1},\cdots,z_{n})=M_{D_{t},\gamma_{t},q}(z_{1},\cdots,z_{n})

(stopped when any zjz_{j} exits DtD_{t}) is a local martingale on chordal/radial SLE probability space. For instance, we can use the identity chart of D,D, and then for [h+,h][h^{+},h^{-}]-differentials MM with conformal dimensions [hq+,hq][h_{q}^{+},h_{q}^{-}] at q,q, we have

Mt(z)=(wt(z))h+(wt(z)¯)h(wt(q))hq+(wt(q)¯)hqM(wt(z)).M_{t}(z)=(w_{t}^{\prime}(z))^{h^{+}}(\overline{w_{t}^{\prime}(z)})^{h^{-}}(w_{t}^{\prime}(q))^{h_{q}^{+}}(\overline{w_{t}^{\prime}(q)})^{h_{q}^{-}}M(w_{t}(z)).

The notion of martingale-observables can be extended naturally to chordal/radial SLE[𝜷]\mathrm{SLE}[\boldsymbol{\beta}] with force points and spins for a general symmetric background charge 𝜷=(𝜷+,𝜷)\boldsymbol{\beta}=(\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}) with the neutrality condition (NCb)(\mathrm{NC}_{b}) and a specific charge aa at p,p, see Theorem 1.5 below. The difference λb(βk+)λb(βk+¯)\lambda_{b}(\beta_{k}^{+})-\lambda_{b}(\overline{\beta_{k}^{+}}) is called the (conformal) spin of 𝜷\boldsymbol{\beta} at qk.q_{k}. Here λb(σ):=σ2/2σb\lambda_{b}(\sigma):={\sigma^{2}}/2-\sigma b is the conformal dimension of charge σ.\sigma.

• SLEs with force points and spins.  We now generalize the standard chordal/radial SLEs to SLE[𝜷]\mathrm{SLE}[\boldsymbol{\beta}] with finitely many force points and spins. Let

(1.5) a=±2/κ,b=a(κ/41)a=\pm\sqrt{2/\kappa},\quad b=a(\kappa/4-1)

so that aa and bb satisfy 2a(a+b)=1.2a(a+b)=1. By definition, for a given symmetric background charge 𝜷\boldsymbol{\beta} on S=DdoubleS=D^{\mathrm{double}} satisfying the neutrality condition (NCb)(\mathrm{NC}_{b}) with a specific charge aa at p,p, the chordal (resp. radial) SLE[𝜷]\mathrm{SLE}[\boldsymbol{\beta}] map gtg_{t} satisfies the chordal Loewner equation (1.4) (resp. the radial Loewner equation (1.2)) driven by the real process ξt:\xi_{t}:

(1.6) dξt=κdBt+λ(t)dt,λ(t)=(λgt1),λ=κξlogZ𝜷ξ,\mathrm{d}\xi_{t}=\sqrt{\kappa}\,\mathrm{d}B_{t}+\lambda(t)\,\mathrm{d}t,\quad\lambda(t)=(\lambda\,\|\,g_{t}^{-1}),\quad\lambda=\kappa\,\partial_{\xi}\log Z_{\boldsymbol{\beta}_{\xi}},

(resp. by the real process θt:\theta_{t}:

(1.7) dθt=κdBt+λ(t)dt,λ(t)=(λgt1),λ=κθlogZ𝜷ζ,ζ=eiθ,ζt=eiθt.)\mathrm{d}\theta_{t}=\sqrt{\kappa}\,\mathrm{d}B_{t}+\lambda(t)\,\mathrm{d}t,\quad\lambda(t)=(\lambda\,\|\,g_{t}^{-1}),\quad\lambda=\kappa\,\partial_{\theta}\log Z_{\boldsymbol{\beta}_{\zeta}},\quad\zeta=\mathrm{e}^{i\theta},\quad\zeta_{t}=\mathrm{e}^{i\theta_{t}}.)

The chordal SLE(κ,𝝆)\mathrm{SLE}(\kappa,\boldsymbol{\rho}) driven by

dξt=κdBt+kρkdtξtqk(t),dqk(t)=2qk(t)ξtdt\mathrm{d}\xi_{t}=\sqrt{\kappa}\,\mathrm{d}B_{t}+\sum_{k}\frac{\rho_{k}\,\mathrm{d}t}{\xi_{t}-q_{k}(t)},\qquad\mathrm{d}q_{k}(t)=\frac{2}{q_{k}(t)-\xi_{t}}\,\mathrm{d}t

has the background charge

𝜷=ap+kβkqk+(2baβ)q,\boldsymbol{\beta}=a\cdot p+\sum_{k}\beta_{k}\cdot q_{k}+(2b-a-\beta)\cdot q,

where βk=ρk/2κ\beta_{k}=\rho_{k}/\sqrt{2\kappa} and β=kβk.\beta=\sum_{k}\beta_{k}. It is known that the chordal SLE(κ,𝝆)\mathrm{SLE}(\kappa,\boldsymbol{\rho}) is almost surely a continuous path, see [26, Theorem 1.3]. The radial SLEη(κ,𝝆)\mathrm{SLE}_{\eta}(\kappa,\boldsymbol{\rho}) driven by ζt=eiθt:\zeta_{t}=\mathrm{e}^{i\theta_{t}}:

(1.8) dθt=κdBt+ηdt+ikρk2cotθtϑk(t)2dt,dqk(t)=qk(t)ζt+qk(t)ζtqk(t)dt,qk=eiϑk\mathrm{d}\theta_{t}=\sqrt{\kappa}\,\mathrm{d}B_{t}+\eta\,\mathrm{d}t+i\sum_{k}\frac{\rho_{k}}{2}\cot\frac{\theta_{t}-\vartheta_{k}(t)}{2}\,\mathrm{d}t,\quad\mathrm{d}q_{k}(t)=q_{k}(t)\frac{\zeta_{t}+q_{k}(t)}{\zeta_{t}-q_{k}(t)}\,\mathrm{d}t,\quad q_{k}=\mathrm{e}^{i\vartheta_{k}}

has the background charge

𝜷=ap+kβkqk+(ba+β+iδ2)q+(ba+βiδ2)q,δ=ηa.\boldsymbol{\beta}=a\cdot p+\sum_{k}\beta_{k}\cdot q_{k}+\Big{(}b-\frac{a+\beta+i\delta}{2}\Big{)}\cdot q+\Big{(}b-\frac{a+\beta-i\delta}{2}\Big{)}\cdot q^{*},\qquad\delta=\eta a.

In [27], Miller and Sheffield identified the radial SLEη(κ,𝝆)\mathrm{SLE}_{\eta}(\kappa,\boldsymbol{\rho}) with the flow lines of a (formal) vector field 𝒪𝜷[σz,σz],\mathcal{O}_{\boldsymbol{\beta}}[\sigma\cdot z,-\sigma\cdot z], where σ\sigma is chosen such that the conformal spin s:=λ+λ=2σb=1.s:=\lambda^{+}-\lambda^{-}=-2\sigma b=1.

We now briefly explain how the chordal SLE[𝜷]\mathrm{SLE}[\boldsymbol{\beta}] process can be produced from the standard chordal SLE(κ)\mathrm{SLE}(\kappa) or SLE[𝜷0]\mathrm{SLE}[\boldsymbol{\beta}^{0}] (𝜷0:=ap+(2ba)q\boldsymbol{\beta}^{0}:=a\cdot p+(2b-a)\cdot q) by the density 𝐄𝒪𝜷0[𝜷𝜷0].\mathbf{E}\,\mathcal{O}_{\boldsymbol{\beta}^{0}}[\boldsymbol{\beta}-\boldsymbol{\beta}^{0}]. By Proposition 14.3 in [KM13] or the special case of Theorem 1.5 below, the ratio of partition functions

M:=Z𝜷Z𝜷0=𝐄𝒪𝜷0[𝜷𝜷0]M:=\frac{Z_{\boldsymbol{\beta}}}{Z_{\boldsymbol{\beta}^{0}}}=\mathbf{E}\,\mathcal{O}_{\boldsymbol{\beta}^{0}}[\boldsymbol{\beta}-\boldsymbol{\beta}^{0}]

is a martingale-observable for SLE[𝜷0]\mathrm{SLE}[\boldsymbol{\beta}^{0}] with

dMtMt=κξ|ξ=ξt(logZ𝜷ξZ𝜷ξ0gt1)dBt.\frac{\mathrm{d}M_{t}}{M_{t}}=\sqrt{\kappa}\,\partial_{\xi}|_{\xi=\xi_{t}}(\log\frac{Z_{\boldsymbol{\beta}_{\xi}}}{Z_{\boldsymbol{\beta}_{\xi}^{0}}}\,\|\,g_{t}^{-1})\,\mathrm{d}B_{t}.

From the fact that Z𝜷ξ0=1Z_{\boldsymbol{\beta}_{\xi}^{0}}=1 in the (,)(\mathbb{H},\infty)-uniformization, it follows that the drift term

(1.9) dξ,MtMt=κξ|ξ=ξt(logZ𝜷ξgt1)dt\frac{\mathrm{d}\langle\xi,M\rangle_{t}}{M_{t}}=\kappa\,\partial_{\xi}\big{|}_{\xi=\xi_{t}}(\log Z_{\boldsymbol{\beta}_{\xi}}\,\|\,g_{t}^{-1})\,\mathrm{d}t

from Girsanov’s theorem corresponds to the drift in (1.6), see [33, Theorem 6].

In Subsection 8.3 we present a certain collection of chordal/radial SLE martingale-observables not by Itô’s calculus but conformal field theory.

Theorem 1.5.

Let 𝛃\boldsymbol{\beta} be a symmetric background charge with the neutrality condition (NCb)(\mathrm{NC}_{b}) and a specific charge aa at pp satisfying (1.5). Suppose 𝒳𝛃\mathcal{X}_{\boldsymbol{\beta}} is a string in the OPE family 𝛃\mathcal{F}_{\boldsymbol{\beta}} of Φ𝛃,\Phi_{\boldsymbol{\beta}}, then the non-random field

M=𝐄𝒳𝜷M=\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}}

is a martingale-observable for chordal/radial SLE[𝛃].\mathrm{SLE}[\boldsymbol{\beta}].

In the physics literature, particular forms of this version appeared in [4, 5, 10, 20, 29]. In [14] Dubédat coupled SLE[𝜷]\mathrm{SLE}[\boldsymbol{\beta}] with Φ𝜷.\Phi_{\boldsymbol{\beta}}.

1.3. Examples of SLE martingale-observables

In Subsection 8.4 we present several examples of radial SLE martingale-observables.

• Radial SLE(0)\mathrm{SLE}(0) observables.  The case κ=0\kappa=0 reveals some aspects of “field Markov property.” Indeed, SLE(0)\mathrm{SLE}(0) curves are hyperbolic geodesics, and martingale-observables (of one variable) are non-random fields FFD,p,qF\equiv F_{D,p,q} with the property

F|Dt=FDt,γt,q.F\big{|}_{D_{t}}=F_{D_{t},\gamma_{t},q}.

In a sense one can think of them as integrals of the motion t{Dt,γt,q}t\mapsto\{D_{t},\gamma_{t},q\} in the corresponding Teichmüller space. The reader is invited to check that

arg((1w)w3/2w),Sw+38(ww)2(14w(1w)2)\arg\big{(}(1-w)w^{-3/2}w^{\prime}\big{)},\qquad S_{w}+\frac{3}{8}\Big{(}\frac{w^{\prime}}{w}\Big{)}^{2}\Big{(}1-\frac{4w}{(1-w)^{2}}\Big{)}

are SLE(0)\mathrm{SLE}(0) observables. Here Sw=Nw12Nw2,Nw=(logw)S_{w}=N_{w}^{\prime}-\frac{1}{2}{N_{w}^{2}},\,N_{w}=(\log w^{\prime})^{\prime} are Schwarzian and pre-Schwarzian derivatives of w,w, respectively. (For the first SLE(0)\mathrm{SLE}(0) observable, consider the radial version of Schramm-Sheffield martingale-observables,

a(arg(1w)2w2(κ41)argww),(a=2/κ),a\Big{(}\arg\frac{(1-w)^{2}}{w}-2\big{(}\frac{\kappa}{4}-1\big{)}\arg\frac{w^{\prime}}{w}\Big{)},\qquad(a=\sqrt{2/\kappa}),

(see the first example in Subsection 8.4) and normalize them so that the limit exists as κ0.\kappa\to 0. For the second SLE(0)\mathrm{SLE}(0) observable, consider the 1-point functions of the Virasoro fields,

c12Sw+h1,2w2w(1w)2+h0,1/2w2w2,\frac{c}{12}S_{w}+h_{1,2}\frac{w^{\prime 2}}{w(1-w)^{2}}+h_{0,1/2}\frac{w^{\prime 2}}{w^{2}},

where the central charge cc and the conformal dimensions h1,2,h0,1/2h_{1,2},h_{0,1/2} are given by

c=(3κ8)(6κ)2κ,h1,2=6κ2κ,h0,1/2=(6κ)(κ2)16κ.c=\frac{(3\kappa-8)(6-\kappa)}{2\kappa},\quad h_{1,2}=\frac{6-\kappa}{2\kappa},\quad h_{0,1/2}=\frac{(6-\kappa)(\kappa-2)}{16\kappa}.

See (5.25).) Recently, in [2] multiple SLE(0)\mathrm{SLE}(0) observables are used to show Peltola and Wang’s theorem ([28]): multiple SLE(0)\mathrm{SLE}(0) curves are contained in the real locus of real rational functions with prescribed real critical points.

• Radial SLE(κ)\mathrm{SLE}(\kappa) observables.  For κ>0,\kappa>0, the usual way to find martingale-observables of a given conformal type is using Itô’s calculus. A couple of well-known important examples are referred to below.

Example (κ=2\kappa=2).

The scalar (i.e., [0,0][0,0]-differential)

M=1|w|2|1w|2=P𝔻(1,w)P𝔻(1,0)=PD(p,z)PD(p,q)M=\frac{1-|w|^{2}}{|1-w|^{2}}=\frac{P_{\mathbb{D}}(1,w)}{P_{\mathbb{D}}(1,0)}=\frac{P_{D}(p,z)}{P_{D}(p,q)}

played an important role in the theory of loop erased random walk (LERW), see [25]. Here PDP_{D} is the Poisson kernel of a domain D.D.

Example (κ=6\kappa=6).

The martingale-observable

Mt(z)=et/41wt(z)3wt(z)6M_{t}(z)=\mathrm{e}^{t/4}\frac{\sqrt[3]{1-w_{t}(z)}}{\sqrt[6]{w_{t}(z)}}

is a scalar with respect to zz and a [18,18][\frac{1}{8},\frac{1}{8}]-differential with respect to q.q. Lawler, Schramm, and Werner applied the optional stopping theorem to the martingale Mt(eiθ)M_{t}(\mathrm{e}^{i\theta}) and estimated the probability that the point eiθ\mathrm{e}^{i\theta} is not swallowed by the SLE6\mathrm{SLE}_{6} hull KtK_{t} at time tt to be

𝐏[eiθKt]e2thqsinθ23,hq=18,(0θ<2π).\mathbf{P}[\,\mathrm{e}^{i\theta}\notin K_{t}\,]\asymp\mathrm{e}^{-2th_{q}}\sqrt[3]{\sin\frac{\theta}{2}},\qquad h_{q}=\frac{1}{8},\qquad(0\leq\theta<2\pi).

The exponent 2hq=1/42h_{q}=1/4 is one of many exponents in [24]. See the example (derivative exponents on the boundary) in Subsection 8.5 with κ=6\kappa=6 and h=0.h=0.

• Lawler-Schramm-Werner’s derivative exponents.  Examples of 1-point rooted vertex observables include Lawler-Schramm-Werner’s derivative exponents ([24]) of radial SLEs on the boundary: given hh

𝐄[|wt(eiθ)|h𝟏{τeiθ>t}]e2hqt(sin2θ2)12aσ,\mathbf{E}[|w^{\prime}_{t}(\mathrm{e}^{i\theta})|^{h}\mathbf{1}_{\{\tau_{\mathrm{e}^{i\theta}}>t\}}]\asymp\mathrm{e}^{-2h_{q}t}\Big{(}\sin^{2}\frac{\theta}{2}\Big{)}^{\frac{1}{2}a\sigma},

where σ\sigma and hqh_{q} are given by

σ=a4(κ4+(κ4)2+16κh),hq=σ28+aσ4.\sigma=\frac{a}{4}\big{(}\kappa-4+\sqrt{(\kappa-4)^{2}+16\kappa h}\big{)},\qquad h_{q}=\frac{\sigma^{2}}{8}+\frac{a\sigma}{4}.

• Friedrich-Werner’s formula.  In the chordal case with κ=8/3\kappa=8/3 (c=0),(c=0), it is well known ([15]) that the nn-point function 𝐄^[T(x1)T(xn)id]\widehat{\mathbf{E}}\,[T(x_{1})\cdots T(x_{n})\,\|\,\mathrm{id}_{\mathbb{H}}\,] of Virasoro field TT𝜷,𝜷=2bqT\equiv T_{\boldsymbol{\beta}},\boldsymbol{\beta}=2b\cdot q coincides with Friedrich-Werner’s function B(x1,,xn):B(x_{1},\cdots,x_{n}):

B(x1,,xn)=limε0𝐏(SLE(8/3) hits all [xj,xj+iε2])ε2n.B(x_{1},\cdots,x_{n})=\lim_{\varepsilon\to 0}\frac{\mathbf{P}(\mathrm{SLE}(8/3)\textrm{ hits all }[x_{j},x_{j}+i\varepsilon\sqrt{2}])}{\varepsilon^{2n}}.

We derive the radial version of this formula. See Theorem 8.3.

• Restriction property.  We also use the one-leg operator Λ\Lambda to present a field theoretic proof of the restriction property of radial SLE(8/3)\mathrm{SLE}(8/3): for all hull K,K,

(1.10) 𝐏(SLE(8/3) avoids K)=|ΨK(1)|λ(ΨK(0))μ,(λ=58,μ=548),\mathbf{P}(\mathrm{SLE}(8/3)\textrm{ avoids }K)=|\Psi_{K}^{\prime}(1)|^{\lambda}(\Psi_{K}^{\prime}(0))^{\mu},\qquad\Big{(}\lambda=\frac{5}{8},\,\mu=\frac{5}{48}\Big{)},

where ΨK\Psi_{K} is the conformal map (𝔻K,0)(𝔻,0)(\mathbb{D}\setminus K,0)\to(\mathbb{D},0) satisfying ΨK(0)>0.\Psi_{K}^{\prime}(0)>0. In particular, we explain the restriction exponents λ\lambda and μ\mu in terms of conformal dimensions of the one-leg operators and their effective versions:

λ=h(Λ):=a22ab=6κ2κ,μ=Hq(Λeff):=a24b2=(κ2)(6κ)8κ,\lambda=h(\Lambda):=\frac{a^{2}}{2}-ab=\frac{6-\kappa}{2\kappa},\qquad\mu=H_{q}(\Lambda^{\mathrm{eff}}):=\frac{a^{2}}{4}-b^{2}=\frac{(\kappa-2)(6-\kappa)}{8\kappa},

where Hq=hq++hqH_{q}=h_{q}^{+}+h_{q}^{-} and

Λeff=𝒫qΛ=V[az+(ba2)q+(ba2)q],(𝒫q=C(b)[bq+bq]).\Lambda^{\mathrm{eff}}=\mathcal{P}_{q}\,\Lambda=V\big{[}a\cdot z+\big{(}b-\frac{a}{2}\big{)}\cdot q+\big{(}b-\frac{a}{2}\big{)}\cdot q^{*}\big{]},\qquad(\mathcal{P}_{q}=C_{(b)}[b\cdot q+b\cdot q^{*}]).

See Subsection 4.4.

1.4. Neumann boundary condition and backward SLEs

In Section 9 we sketch an implementation of a version of conformal field theory with Neumann boundary condition. The Gaussian free field N(z,z0)N(z,z_{0}) in DD with Neumann boundary condition can be constructed from the Gaussian free field Ψ(z)\Psi(z) on S=Ddouble:S=D^{\mathrm{double}}:

N(z,z0)=12(Ψ(z,z0)+Ψ(z,z0)).N(z,z_{0})=\frac{1}{\sqrt{2}}\big{(}\Psi(z,z_{0})+\Psi(z^{*},z_{0}^{*})\big{)}.

For a background charge 𝜷\boldsymbol{\beta} on SS with the neutrality condition (NCb),(\mathrm{NC}_{b}), we introduce the background modification N𝜷N_{\boldsymbol{\beta}} of N.N. We present the connection between the OPE family 𝜷N\mathcal{F}_{\boldsymbol{\beta}}^{N} of N𝜷N_{\boldsymbol{\beta}} and the backward chordal/radial SLE[𝜷].\mathrm{SLE}[\boldsymbol{\beta}].

Suppose that the parameters aa and bb are related to the SLE parameter κ\kappa as

a=±2/κ,b=a(κ/4+1).a=\pm\sqrt{2/\kappa},\qquad b=-a(\kappa/4+1).

Let 𝜷\boldsymbol{\beta} be a symmetric background charge with the neutrality condition (NCb)(\mathrm{NC}_{b}) and a specific charge aa at a marked boundary point pD.p\in\partial D. The backward chordal SLE[𝜷]\mathrm{SLE}[\boldsymbol{\beta}] map ftf_{t} from \mathbb{H} satisfies the equation

(1.11) tft(z)=2ft(z)ξt\partial_{t}f_{t}(z)=-\frac{2}{f_{t}(z)-\xi_{t}}

driven by the real process ξt:\xi_{t}:

dξt=κdBt+λ(t)dt,λ(t)=(λft1),λ=κξlogZ𝜷ξ,\mathrm{d}\xi_{t}=\sqrt{\kappa}\,\mathrm{d}B_{t}+\lambda(t)\,\mathrm{d}t,\quad\lambda(t)=(\lambda\,\|\,f_{t}^{-1}),\quad\lambda=\kappa\,\partial_{\xi}\log Z_{\boldsymbol{\beta}_{\xi}},

where the partition function Z𝜷ξZ_{\boldsymbol{\beta}_{\xi}} is given by Z𝜷ξ=C(ib)[i𝜷ξ]Z_{\boldsymbol{\beta}_{\xi}}=C_{(-ib)}[-i\boldsymbol{\beta}_{\xi}] and 𝜷ξ=𝜷+aξap.\boldsymbol{\beta}_{\xi}=\boldsymbol{\beta}+a\cdot\xi-a\cdot p. Its radial counterpart can be defined in a similar way, see Subsection 9.5

We say a non-random conformal field MM is a martingale-observable for backward chordal/radial SLE[𝜷]\mathrm{SLE}[\boldsymbol{\beta}] if for any z1,,zn,z_{1},\cdots,z_{n}\in\mathbb{H}, the process

Mt(z1,,zn)=(Mft1)(z1,,zn)M_{t}(z_{1},\cdots,z_{n})=(M\,\|\,f_{t}^{-1})(z_{1},\cdots,z_{n})

is a local martingale on backward chordal/radial SLE probability space.

Theorem 1.6.

The correlations of fields in 𝛃N\mathcal{F}_{\boldsymbol{\beta}}^{N} form a collection of martingale-observables for backward chordal/radial SLE[𝛃].\mathrm{SLE}[\boldsymbol{\beta}].

In this theory, the central charge cc is given by c=1+12b2=13+12(κ/8+2/κ)25.c=1+12b^{2}=13+12(\kappa/8+2/\kappa)\geq 25.

2. Coulomb gas correlations

In this section we introduce the Coulomb gas correlations as the (holomorphic) differentials with conformal dimensions λj=σj2/2σjb\lambda_{j}={\sigma_{j}^{2}}/2-\sigma_{j}b at zjz_{j}’s (including infinity) and with values

j<kzj,zk(zjzk)σjσk,(zj^)\prod_{\begin{subarray}{c}j<k\\ z_{j},z_{k}\neq\infty\end{subarray}}(z_{j}-z_{k})^{\sigma_{j}\sigma_{k}},\qquad(z_{j}\in\widehat{\mathbb{C}})

in the identity chart of \mathbb{C} and the chart z1/zz\mapsto-1/z at infinity. After explaining this definition, we prove that under the neutrality condition σj=2b\sum\sigma_{j}=2b the Coulomb gas correlation functions are conformally invariant with respect to the Möbius group Aut(^).\mathrm{Aut}(\widehat{\mathbb{C}}).

2.1. Coulomb gas correlations on the Riemann sphere

Let

𝝈=σjzj,\boldsymbol{\sigma}=\sum\sigma_{j}\cdot z_{j},

where {zj}j=1N\{z_{j}\}_{j=1}^{N} is a finite set of (distinct) points on ^\widehat{\mathbb{C}} and σj\sigma_{j}’s are real numbers, (“charges” at zjz_{j}’s), σj=σzj=𝝈(zj).\sigma_{j}=\sigma_{z_{j}}=\boldsymbol{\sigma}(z_{j}). We can think of 𝝈\boldsymbol{\sigma} as a divisor (a function 𝝈:^\boldsymbol{\sigma}:\widehat{\mathbb{C}}\to\mathbb{R} which takes the value 0 at all points except for finitely many points) or as an atomic measure: 𝝈=σjδzj.\boldsymbol{\sigma}=\sum\sigma_{j}\cdot\delta_{z_{j}}. So

𝝈=σj.\int\boldsymbol{\sigma}=\sum\sigma_{j}.

(Some of σj\sigma_{j}’s can be zero, and in any case σz=0\sigma_{z}=0 if zz is not one of zjz_{j}’s. Sometimes we allow σj\sigma_{j}’s to be complex. We need this case e.g., for the one-leg operators with spin s,s\in\mathbb{C}, see Subsection 8.2.) For a divisor 𝝈\boldsymbol{\sigma} in ,\mathbb{C}, i.e., σ=0,\sigma_{\infty}=0, we define the correlation function C[𝝈]C_{\mathbb{C}}[\boldsymbol{\sigma}] by

(2.1) C[𝝈]C𝝈(𝒛)=j<k(zjzk)σjσk.C_{\mathbb{C}}[\boldsymbol{\sigma}]\equiv C_{\mathbb{C}}^{\boldsymbol{\sigma}}(\boldsymbol{z})=\prod_{j<k}(z_{j}-z_{k})^{\sigma_{j}\sigma_{k}}.

This is a holomorphic function of 𝒛\boldsymbol{z} in distinctN={𝒛=(z1,,zN)N|zjzk if jk}.\mathbb{C}^{N}_{\textrm{distinct}}=\{\boldsymbol{z}=(z_{1},\cdots,z_{N})\in\mathbb{C}^{N}\,|\,z_{j}\neq z_{k}\textrm{ if }j\neq k\}.

Typically it is multivalued except in special cases that all σ\sigma’s are integers. (If they are all even, then the order in the product (2.1) does not matter.) If all σj=1,\sigma_{j}=1, then the correlation function is the Vandermonde determinant.

In general, we need to interpret formulas in terms of single-valued branches. (Sometimes, single-valuedness is referred to as “physically.”) If all zjz_{j}’s are in the real line ,\mathbb{R}, then it is physical. If 𝝈\boldsymbol{\sigma} is symmetric or anti-symmetric with respect to \mathbb{R} in ,\mathbb{C}, then the correlation function is single-valued. See Examples (b) and (c) in Subsection 2.5.

We can extend this definition to divisors on ^\widehat{\mathbb{C}} by simply ignoring the charge at infinity: if 𝝈:^,\boldsymbol{\sigma}:\widehat{\mathbb{C}}\to\mathbb{R}, then C^[𝝈]:=C[𝝈|].C_{\widehat{\mathbb{C}}}[\boldsymbol{\sigma}]:=C_{\mathbb{C}}[\boldsymbol{\sigma}|_{\mathbb{C}}].

2.2. Conformal weights and neutrality condition

Fix a (real) parameter b.b. We say that a divisor 𝝈:^\boldsymbol{\sigma}:\widehat{\mathbb{C}}\to\mathbb{R} satisfies the neutrality condition (NCb\mathrm{NC}_{b}) if

(2.2) 𝝈=2b.\int\boldsymbol{\sigma}=2b.

Note that there is a 1-to-1 correspondence between divisors on ^\widehat{\mathbb{C}} satisfying (NCb\mathrm{NC}_{b}) and arbitrary divisors in :\mathbb{C}: 𝝈𝝈|.\boldsymbol{\sigma}\mapsto\boldsymbol{\sigma}|_{\mathbb{C}}. Let

λb(σ)=σ22σb(σ).\lambda_{b}(\sigma)=\frac{\sigma^{2}}{2}-\sigma b\quad(\sigma\in\mathbb{R}).

Using this function λ:,\lambda:\mathbb{R}\to\mathbb{R}, we define the “weights” or “dimensions” λj\lambda_{j} at zjz_{j} by

λj=λb(σj)σj22σjb.\lambda_{j}=\lambda_{b}(\sigma_{j})\equiv\frac{\sigma_{j}^{2}}{2}-\sigma_{j}b.

It is obvious that λb(σ)=λb(2bσ).\lambda_{b}(\sigma)=\lambda_{b}(2b-\sigma).

2.3. Coulomb gas correlation functions as differentials

Let us recall the definitions of conformal fields and certain transformation laws such as differentials. A local coordinate chart on a Riemann surface MM is a conformal map ϕ:Uϕ(U)\phi:U\to\phi(U)\subset\mathbb{C} on an open subset UU of M.M. By definition, a non-random conformal field ff is an assignment of a (smooth) function (fϕ):ϕ(U)(f\,\|\,\phi):~{}\phi(U)\to\mathbb{C} to each local chart ϕ:Uϕ(U).\phi:U\to\phi(U). A non-random conformal field ff is a differential of weights or (conformal) dimensions [λ,λ][\lambda,\lambda_{*}] if for any two overlapping charts ϕ\phi and ϕ~,\widetilde{\phi}, we have

f=(h)λ(h¯)λf~h,f=(h^{\prime})^{\lambda}(\overline{h^{\prime}})^{\lambda_{*}}\widetilde{f}\circ h,

where h=ϕ~ϕ1:ϕ(UU~)ϕ~(UU~)h=\widetilde{\phi}\circ\phi^{-1}:~{}\phi(U\cap\widetilde{U})\to\widetilde{\phi}(U\cap\widetilde{U}) is the transition map, and ff (resp. f~\widetilde{f}) is the notation for (fϕ),(f\,\|\,\phi), (resp. (fϕ~)(f\,\|\,\widetilde{\phi})). Pre-pre-Schwarzian forms of order (μ,ν)(\mu,\nu) (or PPS(μ,ν)\mathrm{PPS}(\mu,\nu) forms), pre-Schwarzian forms of order μ\mu (or PS(μ)\mathrm{PS}(\mu) forms), and Schwarzian forms of order μ\mu are fields with transformation laws

f=f~h+μlogh+νlogh¯,f=hf~h+μNh,f=(h)2f~h+μSh,f=\widetilde{f}\circ h+\mu\log h^{\prime}+\nu\,\log\overline{h^{\prime}},\qquad f=h^{\prime}\widetilde{f}\circ h+\mu N_{h},\qquad f=(h^{\prime})^{2}\widetilde{f}\circ h+\mu S_{h},

respectively, where

Nh=(logh),Sh=Nh12Nh2N_{h}=(\log h^{\prime})^{\prime},\qquad S_{h}=N_{h}^{\prime}-\frac{1}{2}{N_{h}^{2}}

are pre-Schwarzian and Schwarzian derivatives of h.h. The transformation laws can be extended to the random field: e.g., a field XX is called a [λ,λ][\lambda,\lambda_{*}]-differential if the non-random field z𝐄[X(z)𝒴]z\mapsto\mathbf{E}[X(z)\mathcal{Y}] is a [λ,λ][\lambda,\lambda_{*}]-differential in zz for each 𝒴.\mathcal{Y}.

Let SS be a compact Riemann surface of genus zero, and let 𝝈:S\boldsymbol{\sigma}:S\to\mathbb{R} be a divisor satisfying (NCb\mathrm{NC}_{b}). We define the Coulomb gas correlation function

C[𝝈](CS[𝝈])C𝝈(𝒛)C[\boldsymbol{\sigma}](\equiv C_{S}[\boldsymbol{\sigma}])\equiv C^{\boldsymbol{\sigma}}(\boldsymbol{z})

as a (multivalued) holomorphic differential in zjz_{j}’s such that

  1. (a)

    the conformal dimensions at zjz_{j}’s are the numbers λj\lambda_{j}’s;

  2. (b)

    if S=^,S=\widehat{\mathbb{C}}, then C𝝈(𝒛)C_{\mathbb{C}}^{\boldsymbol{\sigma}}(\boldsymbol{z}) is the value of C𝝈(z)C^{\boldsymbol{\sigma}}(z) in the identity charts at finite zjz_{j}’s and the chart z1/zz\mapsto-1/z at infinity (in the case σ0)\sigma_{\infty}\neq 0).

Alternatively, we can restate (b) as follows: if ϕ:S^\phi:S\to\widehat{\mathbb{C}} is a uniformizing conformal map, then

{C[𝝈]ϕ}=C[𝝈ϕ1].\{C[\boldsymbol{\sigma}]\,\|\,\phi\}=C_{\mathbb{C}}[\boldsymbol{\sigma}\circ\phi^{-1}].

(By convention, ϕ\|\phi refers to the chart 1/ϕ-1/\phi at the pre-image ϕ1()\phi^{-1}(\infty) of infinity.)

Examples.

(a) If 𝝈=2bz,\boldsymbol{\sigma}=2b\cdot z, then C[𝝈]C[\boldsymbol{\sigma}] is a (scalar) function of zz namely, C[𝝈]1.C[\boldsymbol{\sigma}]\equiv 1.

(b) If 𝝈=σz1+(2bσ)z2,\boldsymbol{\sigma}=\sigma\cdot z_{1}+(2b-\sigma)z_{2}, then

C[𝝈]=(z1z2)2λ,λ=λb(σ)=12σ(σ2b)C[\boldsymbol{\sigma}]=(z_{1}-z_{2})^{-2\lambda},\qquad\lambda=\lambda_{b}(\sigma)=\frac{1}{2}\sigma(\sigma-2b)

for finite z1,z2z_{1},z_{2} and C[𝝈]1C[\boldsymbol{\sigma}]\equiv 1 for z2=.z_{2}=\infty. We note that C[𝝈]C[\boldsymbol{\sigma}] is a λ\lambda-differential in z1,z2,z_{1},z_{2}, and a more traditional notation would be

(dz1dz2(z1z2)2)λ.\bigg{(}\frac{dz_{1}\,dz_{2}}{(z_{1}-z_{2})^{2}}\!\bigg{)}^{\lambda}.

Möbius invariance of the bi-differential

dz1dz2(z1z2)2\frac{dz_{1}\,dz_{2}}{(z_{1}-z_{2})^{2}}

is of course well known, and could be used as an alternate way to derive our theorem.

2.4. Möbius invariance

To justify the above definition of Coulomb gas correlation function, we need to verify that the differentials C[𝝈]C[\boldsymbol{\sigma}] are Möbius invariant on ^.\widehat{\mathbb{C}}. Since translation invariance of C[𝝈]C[\boldsymbol{\sigma}] is obvious, we need to verify the invariance of C[𝝈]C[\boldsymbol{\sigma}] under the dilation-rotations τ(z)=az\tau(z)=az and the inversion τ(z)=1/z.\tau(z)=-1/z. First we assume σ=0\sigma_{\infty}=0 and denote

C(𝒛)=j<k(zjzk)σjσk,(𝒛=(z1,,zN))C(\boldsymbol{z})=\prod_{j<k}(z_{j}-z_{k})^{\sigma_{j}\sigma_{k}},\quad(\boldsymbol{z}=(z_{1},\cdots,z_{N}))

(so that C=C[𝝈]C=C_{\mathbb{C}}[\boldsymbol{\sigma}]). For a Möbius map τ,\tau, we also denote τ𝒛=(τz1,,τzN).\tau\boldsymbol{z}=(\tau z_{1},\cdots,\tau z_{N}).

Lemma 2.1.

If 𝛔:\boldsymbol{\sigma}:\mathbb{C}\to\mathbb{R} satisfies the neutrality condition (NCb)(\mathrm{NC}_{b}), and if τ\tau is a Möbius map such that τ(zj),\tau(z_{j})\neq\infty, then

(2.3) C(𝒛)=C(τ𝒛)j(τ(zj))λj.C(\boldsymbol{z})=C(\tau\boldsymbol{z})\prod_{j}\big{(}\tau^{\prime}(z_{j})\big{)}^{\lambda_{j}}.
Proof.

For a dilation-rotation τ(z)=az(a{0}),\tau(z)=az\,(a\in\mathbb{C}\setminus\{0\}), we have

C(τ𝒛)=C(𝒛)aj<kσjσk and j(τ(zj))λj=ajλj,C(\tau\boldsymbol{z})=C(\boldsymbol{z})a^{\sum_{j<k}\sigma_{j}\sigma_{k}}\textrm{ and }\prod_{j}\big{(}\tau^{\prime}(z_{j})\big{)}^{\lambda_{j}}=a^{\sum_{j}\lambda_{j}},

so conformal invariance of C[𝝈]C[\boldsymbol{\sigma}] under the dilation-rotations means that

j<kσjσk+12jσj2=jσjb.\sum_{j<k}\sigma_{j}\sigma_{k}+\frac{1}{2}\sum_{j}\sigma_{j}^{2}=\sum_{j}\sigma_{j}b.

This identity holds by the neutrality condition (NCb)(\mathrm{NC}_{b}) – both sides are equal to 12(jσj)2.\frac{1}{2}(\sum_{j}\sigma_{j})^{2}.

For the inversion τ(z)=1/z,\tau(z)=-1/z, we have

C(τ𝒛)=C(𝒛)j<k(zjzk)σjσk and j(τ(zj))λj=jzj2σjbσj2.C(\tau\boldsymbol{z})=C(\boldsymbol{z})\prod_{j<k}(z_{j}z_{k})^{-\sigma_{j}\sigma_{k}}\textrm{ and }\prod_{j}\big{(}\tau^{\prime}(z_{j})\big{)}^{\lambda_{j}}=\prod_{j}z_{j}^{2\sigma_{j}b-\sigma_{j}^{2}}.

Thus (2.3) reduces to

j<k(zjzk)σjσkjzjσj2=jzj2σjb.\prod_{j<k}(z_{j}z_{k})^{\sigma_{j}\sigma_{k}}\prod_{j}z_{j}^{\sigma_{j}^{2}}=\prod_{j}z_{j}^{2\sigma_{j}b}.

This holds by (NCb)(\mathrm{NC}_{b}) because the exponents of zjz_{j} on both sides coincide:

σj2+kjσjσk=σj2+σj(2bσj)=2σjb.\sigma_{j}^{2}+\sum_{k\neq j}\sigma_{j}\sigma_{k}=\sigma_{j}^{2}+\sigma_{j}(2b-\sigma_{j})=2\sigma_{j}b.

Theorem 2.2.

Under the neutrality condition (NCb)(\mathrm{NC}_{b}), the differentials C[𝛔]C[\boldsymbol{\sigma}] are Möbius invariant on ^.\widehat{\mathbb{C}}.

Proof.

Due to the previous lemma, it remains to check the Möbius invariance of C[𝝈]C[\boldsymbol{\sigma}] on the Riemann sphere in the case that a charge σ\sigma_{\infty} at infinity is non-zero. Let

𝝈=σ+σjzj.\boldsymbol{\sigma}=\sigma_{\infty}\cdot\infty+\sum\sigma_{j}\cdot z_{j}.

We may assume that 𝝈(0)=0\boldsymbol{\sigma}(0)=0 by translation. For the inversion τ(z)=1/z,\tau(z)=-1/z, denote 𝝈~:=𝝈τ1\widetilde{\boldsymbol{\sigma}}:=\boldsymbol{\sigma}\circ\tau^{-1} and C~(𝒛):=C[𝝈~].\widetilde{C}(\boldsymbol{z}):=C_{\mathbb{C}}[\widetilde{\boldsymbol{\sigma}}]. Then 𝝈~=σ0+σj(1/zj)\widetilde{\boldsymbol{\sigma}}=\sigma_{\infty}\cdot 0+\sum\sigma_{j}\cdot(-1/z_{j}) and

(2.4) C~(𝒛)=C(𝒛)j<k(zjzk)σjσkjzjσσj.\widetilde{C}(\boldsymbol{z})=C(\boldsymbol{z})\prod_{j<k}(z_{j}z_{k})^{-\sigma_{j}\sigma_{k}}\prod_{j}z_{j}^{-\sigma_{\infty}\sigma_{j}}.

It follows from C(𝒛)={C[𝝈]id}C(\boldsymbol{z})=\{C[\boldsymbol{\sigma}]\,\|\,\mathrm{id}\} and C~(𝒛)={C[𝝈]τ}\widetilde{C}(\boldsymbol{z})=\{C[\boldsymbol{\sigma}]\,\|\,\tau\} that the conformal invariance of C[𝝈]C[\boldsymbol{\sigma}] under τ\tau means

(2.5) C(𝒛)=C~(𝒛)(h(0))λj(τ(zj))λj=C~(𝒛)jzj2λj,C(\boldsymbol{z})=\widetilde{C}(\boldsymbol{z})\big{(}h^{\prime}(0)\big{)}^{\lambda_{\infty}}\prod_{j}\big{(}\tau^{\prime}(z_{j})\big{)}^{\lambda_{j}}=\widetilde{C}(\boldsymbol{z})\prod_{j}z_{j}^{-2\lambda_{j}},

where h=idh=\mathrm{id} is the transition map between the two charts at infinity. The equations (2.4) – (2.5) reduce to the identity

j<k(zjzk)σjσkjzjσσj+2λj=1.\prod_{j<k}(z_{j}z_{k})^{\sigma_{j}\sigma_{k}}\prod_{j}z_{j}^{\sigma_{\infty}\sigma_{j}+2\lambda_{j}}=1.

This holds by the neutrality condition (NCb)(\mathrm{NC}_{b}) because the exponent of zjz_{j} on the left-hand side is

σj22σjb+σj(σ+kjσk)=0.\sigma_{j}^{2}-2\sigma_{j}b+\sigma_{j}\Big{(}\sigma_{\infty}+\sum_{k\neq j}\sigma_{k}\Big{)}=0.

Finally, for a dilation-rotation τ(z)=az(a0),\tau(z)=az\,(a\neq 0), we have 𝝈~:=𝝈τ1=σ+σj(azj)\widetilde{\boldsymbol{\sigma}}:=\boldsymbol{\sigma}\circ\tau^{-1}=\sigma_{\infty}\cdot\infty+\sum\sigma_{j}\cdot(az_{j}) and

C~(𝒛)=C(𝒛)aj<kσjσk.\widetilde{C}(\boldsymbol{z})=C(\boldsymbol{z})a^{\sum_{j<k}\sigma_{j}\sigma_{k}}.

We need to check that

C(𝒛)=C~(𝒛)(h(0))λj(τ(zj))λj=C~(z)aλ+jλj,C(\boldsymbol{z})=\widetilde{C}(\boldsymbol{z})\big{(}h^{\prime}(0)\big{)}^{\lambda_{\infty}}\prod_{j}\big{(}\tau^{\prime}(z_{j})\big{)}^{\lambda_{j}}=\widetilde{C}(z)a^{-\lambda_{\infty}+\sum_{j}\lambda_{j}},

where h(w)=w/ah(w)=w/a is the transition map between the two charts at infinity. (Indeed, the first chart is z1/zz\mapsto-1/z and the second one is z1/τ(z).z\mapsto-1/\tau(z).) Thus the condition for conformal invariance is

jλj+j<kσjσk=λ.\sum_{j}\lambda_{j}+\sum_{j<k}\sigma_{j}\sigma_{k}=\lambda_{\infty}.

The left-hand side simplifies to λb(jσj)\lambda_{b}(\sum_{j}\sigma_{j}). It follows from the neutrality condition (NCb\mathrm{NC}_{b}) that

λb(jσj)=λb(2bσ)=λb(σ)=λ.\lambda_{b}(\sum_{j}\sigma_{j})=\lambda_{b}(2b-\sigma_{\infty})=\lambda_{b}(\sigma_{\infty})=\lambda_{\infty}.

2.5. Schottky double construction

In this subsection we introduce Coulomb gas correlations for simply-connected domain .\subsetneq\mathbb{C}. They are constructed from those of the Schottky double. We compute them in the \mathbb{H}- and 𝔻\mathbb{D}-uniformizations.

Suppose DD is a simply-connected domain (DD\subsetneq\mathbb{C}). Let D\partial D be its Carathéodory “boundary” (prime ends). Consider the Schottky double S=Ddouble,S=D^{\mathrm{double}}, which equips with the canonical involution ιιD:SS,zz.\iota\equiv\iota_{D}:S\to S,\,z\mapsto z^{*}. For example, we identify ^\widehat{\mathbb{C}} with the Schottky double of \mathbb{H} or that of 𝔻.\mathbb{D}. Then the corresponding involution ι\iota is ι:zz=z¯\iota:z\mapsto z^{*}=\bar{z} for D=D=\mathbb{H} and ι:zz=1/z¯\iota:z\mapsto z^{*}=1/\bar{z} for D=𝔻.D=\mathbb{D}.

For two divisors 𝝈+=σj+zj,\boldsymbol{\sigma}^{+}=\sum\sigma_{j}^{+}\cdot z_{j}, and 𝝈=σjzj\boldsymbol{\sigma}^{-}=\sum\sigma_{j}^{-}\cdot z_{j} (σj+,σj)(\sigma_{j}^{+},\sigma_{j}^{-}\in\mathbb{C}) in DD,D\cup\partial D, we define

𝝈=𝝈++𝝈,𝝈:=σjzj.\boldsymbol{\sigma}=\boldsymbol{\sigma}^{+}+\boldsymbol{\sigma}_{*}^{-},\qquad\boldsymbol{\sigma}_{*}^{-}:=\sum\sigma_{j}^{-}\cdot z_{j}^{*}.

Then 𝝈\boldsymbol{\sigma} is a divisor in S.S. (We may assume that 𝝈\boldsymbol{\sigma}^{-} is a divisor in DD, i.e., σj=0\sigma_{j}^{-}=0 if zjD.z_{j}\in\partial D.) We call (𝝈+,𝝈)(\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}) a double divisor in DD.D\cup\partial D. By definition,

CD(𝝈+,𝝈)(𝒛)(CD[𝝈+,𝝈]):=CS[𝝈].C_{D}^{(\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-})}(\boldsymbol{z})(\equiv C_{D}[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]):=C_{S}[\boldsymbol{\sigma}].

(We often omit the subscripts D,SD,S when there is no danger of confusion.) More precisely, the above definition means

(2.6) {C[𝝈+,𝝈]ϕ}={C[𝝈](ϕ,iϕι)},\{C[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]\,\|\,\phi\}=\{C[\boldsymbol{\sigma}]\,\|\,(\phi,i\circ\phi\circ\iota)\},

where ii is the complex conjugation.

Let λj±=λb(σj±).\lambda_{j}^{\pm}=\lambda_{b}(\sigma_{j}^{\pm}). The following theorem is immediate from Theorem 2.2.

Theorem 2.3.

If 𝛔++𝛔\boldsymbol{\sigma}^{+}+\boldsymbol{\sigma}^{-}_{*} satisfies the neutrality condition (NCb),(\mathrm{NC}_{b}), then C[𝛔+,𝛔]C[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}_{*}] is a well-defined differential with conformal dimensions [λj+,λj][\lambda_{j}^{+},\lambda_{j}^{-}] at zj.z_{j}.

\mathbb{H}-uniformization

Let us consider D=.D=\mathbb{H}. Then its prime end is D=^.\partial D=\widehat{\mathbb{R}}. We may assume that σj=0\sigma_{j}^{-}=0 if zj^.z_{j}\in\widehat{\mathbb{R}}. We now define

C[𝝈+,𝝈]=j<k(zjzk)σj+σk+(z¯jz¯k)σjσkj,k(zjz¯k)σj+σk,C_{\mathbb{H}}[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]=\prod_{j<k}(z_{j}-z_{k})^{\sigma_{j}^{+}\sigma_{k}^{+}}(\bar{z}_{j}-\bar{z}_{k})^{\sigma_{j}^{-}\sigma_{k}^{-}}\prod_{j,k}(z_{j}-\bar{z}_{k})^{\sigma_{j}^{+}\sigma_{k}^{-}},

where the product is taken over finite zjz_{j}’s and zkz_{k}’s, and as always we use the convention 00:=1.0^{0}:=1.

Theorem 2.4.

Under the neutrality condition (NCb),(\mathrm{NC}_{b}), C[𝛔+,𝛔]C_{\mathbb{H}}[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}] is the value of the differential C[𝛔+,𝛔]C[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}] in the identity chart of \mathbb{H} (and the chart z1/zz\mapsto-1/z at infinity).

Remark.

The expression C[𝝈+,𝝈]C_{\mathbb{H}}[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}] of course makes sense without any neutrality condition but we should always reconstruct neutrality by adjusting the charge at infinity. (Recall that there is a 1-to-1 correspondence between divisors on ^\widehat{\mathbb{C}} satisfying (NCb\mathrm{NC}_{b}) and arbitrary divisors in :\mathbb{C}: 𝝈𝝈|.\boldsymbol{\sigma}\mapsto\boldsymbol{\sigma}|_{\mathbb{C}}.)

We leave the proof of the above theorem as a trivial exercise. In the next subsection we state the version of this theorem in the 𝔻\mathbb{D}-uniformization and provide its proof. For a divisor 𝝈+=σj+zj\boldsymbol{\sigma}^{+}=\sum\sigma_{j}^{+}\cdot z_{j} in DD,D\cup\partial D, we define 𝝈+¯=σj+¯zj.\overline{\boldsymbol{\sigma}^{+}}=\sum\overline{\sigma_{j}^{+}}\cdot z_{j}.

Examples.

We have

(a) if 𝝈=𝟎,\boldsymbol{\sigma}^{-}=\boldsymbol{0}, then

C[𝝈+,𝟎]=j<k(zjzk)σj+σk+;C_{\mathbb{H}}[\boldsymbol{\sigma}^{+},\boldsymbol{0}]=\prod_{j<k}(z_{j}-z_{k})^{\sigma_{j}^{+}\sigma_{k}^{+}};

(b) if 𝝈=𝝈+¯,\boldsymbol{\sigma}^{-}=\overline{\boldsymbol{\sigma}^{+}}, then (up to a phase)

C[𝝈+,𝝈+¯]=j<k|(zjzk)σj+σk+(zjz¯k)σj+σk+¯|2Imzj>0(2Imzj)|σj+|2;C_{\mathbb{H}}[\boldsymbol{\sigma}^{+},\overline{\boldsymbol{\sigma}^{+}}]=\prod_{j<k}\Big{|}(z_{j}-z_{k})^{\sigma_{j}^{+}\sigma_{k}^{+}}(z_{j}-\bar{z}_{k})^{\sigma_{j}^{+}\overline{\sigma_{k}^{+}}}\Big{|}^{2}\prod_{\mathrm{Im}\,z_{j}>0}(2\,\mathrm{Im}\,z_{j})^{|\sigma_{j}^{+}|^{2}};

(c) if 𝝈=𝝈+¯,\boldsymbol{\sigma}^{-}=-\,\overline{\boldsymbol{\sigma}^{+}}, then (up to a phase)

C[𝝈+,𝝈+¯]=j<k|(zjzk)σj+σk+(zjz¯k)σj+σk+¯|2Imzj>0(2Imzj)|σj+|2.C_{\mathbb{H}}[\boldsymbol{\sigma}^{+},-\,\overline{\boldsymbol{\sigma}^{+}}]=\prod_{j<k}\Big{|}(z_{j}-z_{k})^{\sigma_{j}^{+}\sigma_{k}^{+}}(z_{j}-\bar{z}_{k})^{-\sigma_{j}^{+}\overline{\sigma_{k}^{+}}}\Big{|}^{2}\prod_{\mathrm{Im}\,z_{j}>0}(2\,\mathrm{Im}\,z_{j})^{-|\sigma_{j}^{+}|^{2}}.

(The products are taken over finite zjz_{j}’s and zkz_{k}’s.)

𝔻\mathbb{D}-uniformization

In the unit disc 𝔻,\mathbb{D}, we define

C𝔻[𝝈+,𝝈]=j<k(zjzk)σj+σk+(z¯jz¯k)σjσkj,k(1zjz¯k)σj+σk.C_{\mathbb{D}}[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]=\prod_{j<k}(z_{j}-z_{k})^{\sigma_{j}^{+}\sigma_{k}^{+}}(\bar{z}_{j}-\bar{z}_{k})^{\sigma_{j}^{-}\sigma_{k}^{-}}\prod_{j,k}(1-z_{j}\bar{z}_{k})^{\sigma_{j}^{+}\sigma_{k}^{-}}.
Theorem 2.5.

Under the neutrality condition (NCb),(\mathrm{NC}_{b}), C𝔻[𝛔+,𝛔]C_{\mathbb{D}}[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}] is the value of the differential C[𝛔+,𝛔]C[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}] in the identity chart of 𝔻.\mathbb{D}.

Proof.

We identify ^\widehat{\mathbb{C}} with the Schottky double of 𝔻.\mathbb{D}. Then the corresponding involution ι\iota is

ι:zz=1z¯.\iota:z\mapsto z^{*}=\frac{1}{\bar{z}}.

A double divisor (𝝈+,𝝈)(\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}) corresponds to a divisor 𝝈\boldsymbol{\sigma} on ^;\widehat{\mathbb{C}}; 𝝈=𝝈++𝝈=(σj+zj+σjzj).\boldsymbol{\sigma}=\boldsymbol{\sigma}^{+}+\boldsymbol{\sigma}^{-}_{*}=\sum(\sigma_{j}^{+}\cdot z_{j}+\sigma_{j}^{-}\cdot z_{j}^{*}). If one of the zjz_{j}’s is 0, we call it z0z_{0} (so that z0=0z_{0}=0). Denote

C~={C[𝝈+,𝝈]id𝔻}.\widetilde{C}=\{C[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]\,\|\,\mathrm{id}_{\mathbb{D}}\}.

We need to show that C~=C𝔻.\widetilde{C}=C_{\mathbb{D}}. By definition (2.6),

C~={C[𝝈](id,ι¯)},\widetilde{C}=\{C[\boldsymbol{\sigma}]\,\|\,(\mathrm{id},\bar{\iota})\},

where we use the chart ι¯:z1/z\bar{\iota}:z\mapsto 1/z at the nodes zjz_{j}^{*} including z0=.z_{0}^{*}=\infty. It follows that

C:=C[𝝈]=C~j>0(z¯j)2λj.C:=C_{\mathbb{C}}[\boldsymbol{\sigma}]=\widetilde{C}\prod_{j>0}(-\bar{z}_{j})^{2\lambda_{j}^{-}}.

Indeed, CC is the value of C[𝝈]C[\boldsymbol{\sigma}] in the identity chart of \mathbb{C} and the chart ι¯\bar{\iota} at infinity, so

C=C~j>0(h(zj))λj,C=\widetilde{C}\prod_{j>0}\big{(}h^{\prime}(z_{j}^{*})\big{)}^{\lambda_{j}^{-}},

where h=ι¯h=\bar{\iota} is the transition map. Recall the expression for C:C_{\mathbb{C}}:

C=j<k(zjzk)σj+σk+0<j<k(1/z¯j1/z¯k)σjσkj,k;k0(zj1/z¯k)σj+σk.C=\prod_{j<k}(z_{j}-z_{k})^{\sigma_{j}^{+}\sigma_{k}^{+}}\prod_{0<j<k}(1/\bar{z}_{j}-1/\bar{z}_{k})^{\sigma_{j}^{-}\sigma_{k}^{-}}\prod_{j,k;k\neq 0}(z_{j}-1/\bar{z}_{k})^{\sigma_{j}^{+}\sigma_{k}^{-}}.

Let us rewrite this as a fraction C=N/DC=N/D with

N=j<k(zjzk)σj+σk+0<j<k(z¯jz¯k)σjσkj,k(1zjz¯k)σj+σk,N=\prod_{j<k}(z_{j}-z_{k})^{\sigma_{j}^{+}\sigma_{k}^{+}}\prod_{0<j<k}(\bar{z}_{j}-\bar{z}_{k})^{\sigma_{j}^{-}\sigma_{k}^{-}}\prod_{j,k}(1-z_{j}\bar{z}_{k})^{\sigma_{j}^{+}\sigma_{k}^{-}},

and

D=0<j<k(z¯jz¯k)σjσkj,k;k0(z¯k)σj+σk.D=\prod_{0<j<k}(-\bar{z}_{j}\bar{z}_{k})^{\sigma_{j}^{-}\sigma_{k}^{-}}\prod_{j,k;k\neq 0}(-\bar{z}_{k})^{\sigma_{j}^{+}\sigma_{k}^{-}}.

Comparing C𝔻C_{\mathbb{D}} to N,N, we have

C𝔻\displaystyle C_{\mathbb{D}} =Nj=0,k>0(z¯jz¯k)σjσk=Nk>0(z¯k)σ0σk\displaystyle=N\prod_{j=0,k>0}(\bar{z}_{j}-\bar{z}_{k})^{\sigma_{j}^{-}\sigma_{k}^{-}}=N\prod_{k>0}(-\bar{z}_{k})^{\sigma_{0}^{-}\sigma_{k}^{-}}
=CDk>0(z¯k)σ0σk=C~Dj>0(z¯j)2λjk>0(z¯k)σ0σk.\displaystyle=CD\prod_{k>0}(-\bar{z}_{k})^{\sigma_{0}^{-}\sigma_{k}^{-}}=\widetilde{C}D\prod_{j>0}(-\bar{z}_{j})^{2\lambda_{j}^{-}}\prod_{k>0}(-\bar{z}_{k})^{\sigma_{0}^{-}\sigma_{k}^{-}}.

To verify C𝔻=C~,C_{\mathbb{D}}=\widetilde{C}, it remains to show that

j>0(z¯j)2λjj>0(z¯j)σ0σj0<j<k(z¯jz¯k)σjσkj,k;k0(z¯k)σj+σk=1.\prod_{j>0}(-\bar{z}_{j})^{2\lambda_{j}^{-}}\prod_{j>0}(-\bar{z}_{j})^{\sigma_{0}^{-}\sigma_{j}^{-}}\prod_{0<j<k}(-\bar{z}_{j}\bar{z}_{k})^{\sigma_{j}^{-}\sigma_{k}^{-}}\prod_{j,k;k\neq 0}(-\bar{z}_{k})^{\sigma_{j}^{+}\sigma_{k}^{-}}=1.

The exponent of z¯j(j0)\bar{z}_{j}\,(j\neq 0) on the left-hand side is

2λj+σ0σj\displaystyle 2\lambda_{j}^{-}+\sigma_{0}^{-}\sigma_{j}^{-} +kj,0σjσk+kσk+σj=2λj+σj(kjσk+kσk+)\displaystyle+\sum_{k\neq j,0}\sigma_{j}^{-}\sigma_{k}^{-}+\sum_{k}\sigma_{k}^{+}\sigma_{j}^{-}=2\lambda_{j}^{-}+\sigma_{j}^{-}\Big{(}\sum_{k\neq j}\sigma_{k}^{-}+\sum_{k}\sigma_{k}^{+}\Big{)}
=2λj+σj(2bσj)=2λj2λb(σj)=0.\displaystyle=2\lambda_{j}^{-}+\sigma_{j}^{-}(2b-\sigma_{j}^{-})=2\lambda_{j}^{-}-2\lambda_{b}(\sigma_{j}^{-})=0.

For the exterior Δ\Delta of the unit disc, we leave it to the readers to check that the Coulomb gas correlation functions in the Δ\Delta-uniformization have the same expression as in the 𝔻\mathbb{D}-uniformization but one must disregard the node at infinity (as in the \mathbb{H}-uniformization).

3. Conformal field theory of Gaussian free field

In this section we review a version of conformal field theory with central charge c=1c=1 (i.e., b=0b=0) both in a simply-connected domain and on the Riemann sphere implemented in [18, 17] and present their connection in the context of Schottky double construction. Coulomb gas correlations in the case b=0b=0 are represented as correlations of (formal) multi-vertex fields constructed from the Gaussian free field through Wick’s calculus. In Sections 4 – 5, we extend multi-vertex fields to the case b0b\neq 0 and interpret them as the OPE exponentials of background charge modifications of bosonic fields.

3.1. Bosonic field and its Wick’s exponentials

The chiral bosonic fields are described as the holomorphic part or anti-holomorphic part of the Gaussian free field in the physics literature.

• Gaussian free field.  The Gaussian free field Φ\Phi in a planar domain DD with Dirichlet boundary condition is an isometry Φ:(D)L2(Ω,𝐏)\Phi:\mathcal{E}(D)\to L^{2}(\Omega,\mathbf{P}) from the Dirichlet energy space (D)\mathcal{E}(D) such that the image consists of centered Gaussian random variables. Here (Ω,𝐏)(\Omega,\mathbf{P}) is a probability space and (D)\mathcal{E}(D) is the completion of smooth functions with compact supports in DD with respect to the norm

f2=2G(ζ,z)f(ζ)f(z)¯dA(ζ)dA(z),\|f\|^{2}_{\mathcal{E}}=\iint 2G(\zeta,z)\,f(\zeta)\,\overline{f(z)}~{}\mathrm{d}A(\zeta)\,\mathrm{d}A(z),

where AA is the (normalized) area measure and GGDG\equiv G_{D} is the Dirichlet Green’s function for D.D. In the upper half-plane, we have

G(ζ,z)=log|ζz¯ζz|,G_{\mathbb{H}}(\zeta,z)=\log\Big{|}\frac{\zeta-\bar{z}}{\zeta-z}\Big{|},

where G(ζ,)=0.G(\zeta,\infty)=0. In the unit disc, we have

G𝔻(ζ,z)=log|1ζz¯ζz|.G_{\mathbb{D}}(\zeta,z)=\log\Big{|}\frac{1-\zeta\bar{z}}{\zeta-z}\Big{|}.

The same formula holds in the exterior Δ\Delta of the unit disc, but GΔ(ζ,)=log|ζ|.G_{\Delta}(\zeta,\infty)=\log|\zeta|.

The Gaussian free field can be viewed as a Fock space field with the nn-point correlation function

𝐄[Φ(z1)Φ(zn)]=k2G(zik,zjk),\mathbf{E}[\Phi(z_{1})\cdots\Phi(z_{n})]=\sum\prod_{k}2G(z_{i_{k}},z_{j_{k}}),

where the sum is over all partitions of the set {1,,n}\{1,\cdots,n\} into disjoint pairs {ik,jk}.\{i_{k},j_{k}\}. This correlation function is a unique continuous function on DdistinctnD^{n}_{\mathrm{distinct}} such that

𝐄[Φ(f1)Φ(fn)]=f1(z1)fn(zn)𝐄[Φ(z1)Φ(zn)]dA(z1)dA(zn)\mathbf{E}[\Phi(f_{1})\cdots\Phi(f_{n})]=\int f_{1}(z_{1})\cdots f_{n}(z_{n})~{}\mathbf{E}[\Phi(z_{1})\cdots\Phi(z_{n})]\,\mathrm{d}A(z_{1})\cdots\mathrm{d}A(z_{n})

for all test functions fjf_{j} with disjoint supports.

• Chiral bosonic fields.  We write J:=ΦJ:=\partial\Phi for the current field. The chiral boson

Φ+(z,z0)=z0zJ(ζ)dζ\Phi^{+}(z,z_{0})=\int_{z_{0}}^{z}J(\zeta)\,\mathrm{d}\zeta

is a well-defined, “multivalued”, path dependent, “generalized” centered Gaussian. More precisely,

Φ+(z,z0)={Φ+(γ)=γJ(ζ)dζ},\Phi^{+}(z,z_{0})=\Big{\{}\Phi^{+}(\gamma)=\int_{\gamma}J(\zeta)\,\mathrm{d}\zeta\Big{\}},

where γ\gamma is a curve from z0z_{0} to z.z. Then the values of Φ+\Phi^{+} are multivalued correlation functionals in the complement of the curve. For example, we have

(3.1) 𝐄[Φ+(z,z0)Φ(z1)]=2(G+(z,z1)G+(z0,z1)),\mathbf{E}[\Phi^{+}(z,z_{0})\Phi(z_{1})]=2(G^{+}(z,z_{1})-G^{+}(z_{0},z_{1})),

where G+G^{+} is the complex Dirichlet Green’s function, 2G+(z,z1)=G(z,z1)+iG~(z,z1).2G^{+}(z,z_{1})=G(z,z_{1})+i\widetilde{G}(z,z_{1}). Here G~\widetilde{G} is the harmonic conjugate of G.G. The multivalued function zG+(z,z1)z\mapsto G^{+}(z,z_{1}) is defined up to constants. Sometimes we work with a uniformization w:(D,q)(𝔻,0).w:(D,q)\to(\mathbb{D},0). In such a case, it is convenient to choose the constant so that

G+(z,z1)=12log1w(z)w(z1)¯w(z)w(z1).G^{+}(z,z_{1})=\frac{1}{2}\log\frac{1-w(z)\overline{w(z_{1})}}{w(z)-w(z_{1})}.

We write

Φ(z,z0)=Φ+(z,z0)¯.\Phi^{-}(z,z_{0})=\overline{\Phi^{+}(z,z_{0})}.

Then we have

Φ(z)Φ(z0)=Φ+(z,z0)+Φ(z,z0).\Phi(z)-\Phi(z_{0})=\Phi^{+}(z,z_{0})+\Phi^{-}(z,z_{0}).

• Wick’s calculus.  For centered jointly Gaussian random variables ξjk,\xi_{jk}, (1jl,1kmj),(1\leq j\leq l,1\leq k\leq m_{j}), let Xj=ξj1ξjmjX_{j}=\xi_{j1}\odot\cdots\odot\xi_{jm_{j}} be Wick’s product of ξjk,(1kmj).\xi_{jk},(1\leq k\leq m_{j}). Then by Wick’s formula we have

(3.2) X1Xl=γ{v,v}𝐄[ξvξv]v′′ξv′′,X_{1}\cdots X_{l}=\sum_{\gamma}\prod_{\{v,v^{\prime}\}}\mathbf{E}[\xi_{v}\xi_{v^{\prime}}]~{}\underset{v^{\prime\prime}}{\textstyle\bigodot}\xi_{v^{\prime\prime}},

where the sum is taken over all Feynman diagrams labeled by ξjk\xi_{jk}’s without edges joining any ξjk1\xi_{jk_{1}} and ξjk2.\xi_{jk_{2}}. Recall that the Feynman diagram labeled by ξ1,,ξn\xi_{1},\cdots,\xi_{n} is a graph with vertices 1,,n1,\cdots,n such that edges {v,v}\{v,v^{\prime}\} have no common endpoints. (Such edges {v,v}\{v,v^{\prime}\} are called “Wick’s contractions.”) We denote the unpaired vertices by v′′.v^{\prime\prime}. For generalized Gaussians or Fock space fields like ξjk=Φ(zjk)\xi_{jk}=\Phi(z_{jk}) or Φ+(ζjk,zjk),\Phi^{+}(\zeta_{jk},z_{jk}), we set the tensor product X1XlX_{1}\cdots X_{l} with disjoint sets S(Xj)S(X_{j}) to be (3.2). For example,

Φ2(ζ)Φ2(z)=Φ2(ζ)Φ2(z)+4𝐄[Φ(ζ)Φ(z)]Φ(ζ)Φ(z)+2(𝐄[Φ(ζ)Φ(z)])2,\Phi^{\odot 2}(\zeta)\Phi^{\odot 2}(z)=\Phi^{\odot 2}(\zeta)\odot\Phi^{\odot 2}(z)+4\,\mathbf{E}[\Phi(\zeta)\Phi(z)]\Phi(\zeta)\odot\Phi(z)+2(\mathbf{E}[\Phi(\zeta)\Phi(z)])^{2},

where each of terms on the right hand-side comes from 0,1,0,1, or 2 contractions, respectively. Wick’s exponentials of (generalized) Gaussian ξ\xi is defined by

eξ:=n=0ξnn!.\mathrm{e}^{\odot\xi}:=\sum_{n=0}^{\infty}\frac{\xi^{\odot n}}{n!}.

For centered jointly (generalized) Gaussians ξj\xi_{j}’s we have

eα1ξ1eαnξn=ej<kαjαk𝐄[ξjξk]ejαjξj.\mathrm{e}^{\odot\alpha_{1}\xi_{1}}\cdots\mathrm{e}^{\odot\alpha_{n}\xi_{n}}=\mathrm{e}^{\sum_{j<k}\alpha_{j}\alpha_{k}\mathbf{E}[\xi_{j}\xi_{k}]}\mathrm{e}^{\odot\sum_{j}\alpha_{j}\xi_{j}}.

3.2. Formal bosonic fields

To make the computation easy, it is convenient to consider a representation

Φ+(z,z0)=Φ+(z)Φ+(z0)\Phi^{+}(z,z_{0})=\Phi^{+}(z)-\Phi^{+}(z_{0})

such that

Φ=Φ++Φ,Φ=Φ+¯.\Phi=\Phi^{+}+\Phi^{-},\qquad\Phi^{-}=\overline{\Phi^{+}}.

It is not possible to define such Φ±\Phi^{\pm} in a conformally invariant way, but linear combinations

Φ[𝝈+,𝝈]:=σj+Φ+(zj)σjΦ(zj)\Phi[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]:=\sum\sigma_{j}^{+}\Phi^{+}(z_{j})-\sigma_{j}^{-}\Phi^{-}(z_{j})

satisfying the neutrality condition are well-defined as Fock space fields.

Lemma 3.1.

If a double divisor (𝛔+,𝛔)(\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}) satisfies (NC0)(\mathrm{NC}_{0}), i.e.,

𝝈++𝝈=0,\int\boldsymbol{\sigma}^{+}+\boldsymbol{\sigma}^{-}=0,

then the formal bosonic field Φ[𝛔+,𝛔]\Phi[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}] can be represented as a linear combination of well-defined Fock space fields.

Proof.

Let us choose any point z0Dz_{0}\in D. (It can be one of zjz_{j}’s.) Then

Φ[𝝈+,𝝈]=Φ+(z0)𝝈+Φ(z0)𝝈+σj+Φ+(zj,z0)σjΦ(zj,z0).\Phi[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]=\Phi^{+}(z_{0})\int\boldsymbol{\sigma}^{+}-\Phi^{-}(z_{0})\int\boldsymbol{\sigma}^{-}+\sum\sigma_{j}^{+}\Phi^{+}(z_{j},z_{0})-\sigma_{j}^{-}\Phi^{-}(z_{j},z_{0}).

Under the neutrality condition, the first two terms on the right-hand side become the Fock space correlation functional:

Φ+(z0)𝝈+Φ(z0)𝝈=Φ(z0)𝝈+.\Phi^{+}(z_{0})\int\boldsymbol{\sigma}^{+}-\Phi^{-}(z_{0})\int\boldsymbol{\sigma}^{-}=\Phi(z_{0})\int\boldsymbol{\sigma}^{+}.

Remark.

The representation in the lemma is not unique, of course, but it is “unique” in the sense of “multivalued” fields. For example, Φ[1z11z2,𝟎]=Φ+(z1,z0)Φ+(z2,z0)\Phi[1\cdot z_{1}-1\cdot z_{2},\boldsymbol{0}]=\Phi^{+}(z_{1},z_{0})-\Phi^{+}(z_{2},z_{0}) is “independent” of z0.z_{0}. If we specify the curves in the definition of bi-vertex fields with two different choices of z0,z_{0}, then the difference is an integral over a loop. If we choose z0=z2,z_{0}=z_{2}, then Φ[1z11z2,𝟎]=Φ+(z1,z2)Φ+(z2,z2),\Phi[1\cdot z_{1}-1\cdot z_{2},\boldsymbol{0}]=\Phi^{+}(z_{1},z_{2})-\Phi^{+}(z_{2},z_{2}), where zΦ+(z,z)z\mapsto\Phi^{+}(z,z) is a “monodromy field:”

Φ+(z,z)={γJ(ζ)dζ:γ is a loop rooted at z}.\Phi^{+}(z,z)=\Big{\{}\int_{\gamma}J(\zeta)\,\mathrm{d}\zeta\,:\,\gamma\textrm{ is a loop rooted at }z\Big{\}}.

3.3. Correlations of formal fields in the unit disc and the upper half-plane

We define Φ±Φ𝔻±\Phi^{\pm}\equiv\Phi^{\pm}_{\mathbb{D}} (Φ=Φ+¯)(\Phi^{-}=\overline{\Phi^{+}}) in 𝔻\mathbb{D} as centered Gaussian (formal) fields with “formal” correlations

(3.3) 𝐄[Φ+(z1)Φ+(z2)]=log1z1z2,𝐄[Φ+(z1)Φ(z2)]=log(1z1z¯2).\mathbf{E}[\Phi^{+}(z_{1})\Phi^{+}(z_{2})]=\log\frac{1}{z_{1}-z_{2}},\qquad\mathbf{E}[\Phi^{+}(z_{1})\Phi^{-}(z_{2})]=\log(1-z_{1}\bar{z}_{2}).

Note that these formal correlations have no Aut(𝔻)\mathrm{Aut}(\mathbb{D})-invariance and depend on the order of particles. It is easy to verify that

𝐄[Φ(z1)Φ(z2)]=𝐄[Φ+(z1)+Φ(z1)][Φ+(z2)+Φ(z2)],\mathbf{E}[\Phi(z_{1})\Phi(z_{2})]=\mathbf{E}[\Phi^{+}(z_{1})+\Phi^{-}(z_{1})][\Phi^{+}(z_{2})+\Phi^{-}(z_{2})],

where 𝐄\mathbf{E} on the left-hand side has the usual meaning, but we use “formal” correlations on the right-hand side. The same interpretation is applied for 𝐄[Φ+(z1,z2)Φ+(z3,z4)]\mathbf{E}[\Phi^{+}(z_{1},z_{2})\Phi^{+}(z_{3},z_{4})], etc. If a double divisor (𝝈+,𝝈)(\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}) in DD satisfies the neutrality condition (NC0),(\mathrm{NC}_{0}), then we can compute correlations of Φ[𝝈+,𝝈]\Phi[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}] (or other functionals involving Φ[𝝈+,𝝈],\Phi[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}], e.g., Wick’s exponentials of Φ[𝝈+,𝝈]\Phi[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]) with various Fock space fields by applying Wick’s calculus to our formal fields in 𝔻.\mathbb{D}. In the case of an arbitrary simply-connected domain D,D, we fix a conformal map w:D𝔻w:D\to\mathbb{D} and define Φ±(z)=Φ𝔻±(w(z))\Phi^{\pm}(z)=\Phi^{\pm}_{\mathbb{D}}(w(z)) so the correlations are

𝐄[Φ+(z1)Φ+(z2)]=log1w(z1)w(z2),𝐄[Φ+(z1)Φ(z2)]=log(1w(z1)w(z2)¯).\mathbf{E}[\Phi^{+}(z_{1})\Phi^{+}(z_{2})]=\log\frac{1}{w(z_{1})-w(z_{2})},\quad\mathbf{E}[\Phi^{+}(z_{1})\Phi^{-}(z_{2})]=\log(1-w(z_{1})\overline{w(z_{2})}).

The “formal” correlations depend on the choice of the conformal map but this dependence disappears under the neutrality condition. In particular, we can use this method to introduce formal bosonic fields in \mathbb{H} but it is more convenient to define Φ±Φ±\Phi^{\pm}\equiv\Phi^{\pm}_{\mathbb{H}} in \mathbb{H} as follows:

𝐄[Φ+(z1)Φ+(z2)]=log1z1z2,𝐄[Φ+(z1)Φ(z2)]=log(z1z¯2)\mathbf{E}[\Phi^{+}(z_{1})\Phi^{+}(z_{2})]=\log\frac{1}{z_{1}-z_{2}},\qquad\mathbf{E}[\Phi^{+}(z_{1})\Phi^{-}(z_{2})]=\log(z_{1}-\bar{z}_{2})

for finite zjz_{j}’s and set Φ+()=0.\Phi^{+}(\infty)=0.

3.4. OPE exponentials and multi-vertex fields

The OPE product XYX*Y of two fields XX and YY is a generic notation for a coefficient in the operator product expansion X(ζ)Y(z)X(\zeta)Y(z) as ζz,\zeta\to z, i.e., expansion with respect to a chart independent asymptotic scale. Typically (but not always) we use * for the coefficient of the first non-diverging term. So, the OPE product XYX*Y of non-chiral fields is obtained by subtracting all divergent terms in the operator product expansion X(ζ)Y(z)X(\zeta)Y(z) and then taking the limit as ζz.\zeta\to z. This is the case in notation Φ2:=ΦΦ,Φn=ΦΦn1\Phi^{*2}:=\Phi*\Phi,\Phi^{*n}=\Phi*\Phi^{*n-1} in the definition of OPE powers of Φ.\Phi. Let c(z),zDc(z),z\in D denote the logarithm of conformal radius of D,D, i.e.,

c(z)=u(z,z),u(ζ,z):=GD(ζ,z)+log|ζz|.c(z)=u(z,z),\quad u(\zeta,z):=G_{D}(\zeta,z)+\log|\zeta-z|.

Then by Wick’s calculus,

Φ(ζ)Φ(z)=log1|ζz|2+Φ2(z)+o(1),Φ2=2c+Φ2\Phi(\zeta)\Phi(z)=\log\frac{1}{|\zeta-z|^{2}}+\Phi^{*2}(z)+o(1),\quad\Phi^{*2}=2c+\Phi^{\odot 2}

as ζz,ζz.\zeta\to z,\zeta\neq z. In general, unlike Wick’s multiplication, the OPE multiplication is neither commutative nor associative. For example, if ff is a non-random holomorphic function and if X,YX,Y are holomorphic fields, then fX=Xf=fX,f*X=X*f=fX, and (fX)YX(fY).(fX)*Y\neq X*(fY).

We define OPE exponentials of the Gaussian free field by

𝒱(σ):=eiσΦ=n=0(iσ)nn!Φn.\mathcal{V}^{(\sigma)}:=\mathrm{e}^{*i\sigma\Phi}=\sum_{n=0}^{\infty}\frac{(i\sigma)^{n}}{n!}\,\Phi^{*n}.

Then we have

𝒱(σ)(z)=C(z)σ2eiσΦ(z),C(z)=ec(z),\mathcal{V}^{(\sigma)}(z)=C(z)^{-\sigma^{2}}\mathrm{e}^{\odot i\sigma\Phi(z)},\qquad C(z)=\mathrm{e}^{c(z)},

see [18, Proposition 3.3]. More generally, OPE exponentials (non-chiral multi-vertex fields) 𝒱[𝝈]\mathcal{V}[\boldsymbol{\sigma}] of iΦ[𝝈,𝝈]=iσjΦ(zj)i\Phi[\boldsymbol{\sigma},-\boldsymbol{\sigma}]=\sum i\sigma_{j}\Phi(z_{j}) can be defined in a similar way

𝒱[𝝈]:=eiΦ[𝝈,𝝈]\mathcal{V}[\boldsymbol{\sigma}]:=\mathrm{e}^{*i\Phi[\boldsymbol{\sigma},-\boldsymbol{\sigma}]}

and be computed as

𝒱[𝝈]=𝒞[𝝈]eiΦ[𝝈,𝝈],\mathcal{V}[\boldsymbol{\sigma}]=\mathcal{C}[\boldsymbol{\sigma}]\mathrm{e}^{\odot i\Phi[\boldsymbol{\sigma},-\boldsymbol{\sigma}]},

where 𝒞[𝝈]=C[𝝈,𝝈].\mathcal{C}[\boldsymbol{\sigma}]=C[\boldsymbol{\sigma},-\boldsymbol{\sigma}]. See [17, Section 5] for more details.

• Multi-vertex fields.  We now extend the OPE exponentials or multi-vertex fields to chiral fields by stating the definition of the chiral multi-vertex field in terms of Coulomb gas correlation function and Wick’s exponential. Suppose a double divisor (𝝈+,𝝈)(\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}) in DD satisfies the neutrality condition (NC0).(\mathrm{NC}_{0}). Then the Wick’s exponential

V[𝝈+,𝝈]:=eiΦ[𝝈+,𝝈]eiσj+Φ+(zj)σjΦ(zj)V^{\odot}[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]:=\mathrm{e}^{\odot i\Phi[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]}\equiv\mathrm{e}^{\odot i\sum\sigma_{j}^{+}\Phi^{+}(z_{j})-\sigma_{j}^{-}\Phi^{-}(z_{j})}

is a well-defined Fock space functional; V[𝝈+,𝝈]V^{\odot}[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}] is a scalar, i.e., (0,0)(0,0)-differential and

𝐄V[𝝈+,𝝈]=1.\mathbf{E}\,V^{\odot}[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]=1.

We define the multi-vertex field V[𝝈+,𝝈]V[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}] by

V[𝝈+,𝝈]=C[𝝈+,𝝈]V[𝝈+,𝝈].V[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]=C[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]V^{\odot}[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}].

Thus a multi-vertex field consists of two parts – its expectation given by the correlation differential C[𝝈+,𝝈],C[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}], and the Wick exponential V[𝝈+,𝝈].V^{\odot}[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]. The conformal dimensions of V[𝝈+,𝝈]V[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}] at zjz_{j}’s are

λj+=(σj+)22,λj=(σj)22.\lambda_{j}^{+}=\frac{(\sigma_{j}^{+})^{2}}{2},\qquad\lambda_{j}^{-}=\frac{(\sigma_{j}^{-})^{2}}{2}.

(Recall that b=0b=0 in this section.)

• Formal representation of multi-vertex fields.  For a given conformal map w:D𝔻,w:D\to\mathbb{D}, we formally set

Vσ(z)=(w(z))σ2/2eiσΦ+(z),V¯σ(z)=(w(z)¯)σ2/2eiσΦ(z).V^{\sigma}(z)=(w^{\prime}(z))^{\sigma^{2}/2}\mathrm{e}^{\odot i\sigma\Phi^{+}(z)},\qquad\bar{V}^{\sigma}(z)=(\overline{w^{\prime}(z)})^{\sigma^{2}/2}\mathrm{e}^{\odot-i\sigma\Phi^{-}(z)}.

Then under the neutrality condition (NC0),(\mathrm{NC}_{0}),

𝐄V[𝝈+,𝝈]=𝐄Vσ1+(z1)V¯σ1(z1)Vσn+(zn)V¯σn(zn),\mathbf{E}\,V[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]=\mathbf{E}\,V^{\sigma_{1}^{+}}(z_{1})\bar{V}^{\sigma_{1}^{-}}(z_{1})\cdots V^{\sigma_{n}^{+}}(z_{n})\bar{V}^{\sigma_{n}^{-}}(z_{n}),

where the left-hand side is C[𝝈+,𝝈]C[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}] by the definition above and the right-hand side is computed by Wick’s calculus from formal correlations in 𝔻.\mathbb{D}. The result does not depend on the choice of the conformal map w,w, and is the same if we use formal correlations in \mathbb{H} or Δ.\Delta.

3.5. Conformal field theory on the Riemann sphere

In this subsection we define the Gaussian free field on ^\widehat{\mathbb{C}} both as a Gaussian field indexed by the energy space and as a bi-variant Fock space field. After introducing formal bosonic fields, we define multi-vertex fields on ^.\widehat{\mathbb{C}}.

• Energy space.  For a compact Riemann surface SS of genus zero, we denote 𝒲(S)(0)(S)=W1,2(S)/\mathcal{W}(S)\equiv\mathcal{E}^{(0)}(S)=W^{1,2}(S)/\mathbb{C} with the scalar product

(f,g)=Sdfdg¯.(f,g)_{\nabla}=\int_{S}\mathrm{d}f\wedge\overline{\mathrm{d}g}.

The Dirichlet energy space, (S)=(1,1)(S)\mathcal{E}(S)=\mathcal{E}^{(1,1)}(S) is the Hilbert space defined as the completion of the smooth (1,1)(1,1) forms with respect to the scalar product such that the Laplacian operator

¯:𝒲(S)(S)\partial{\bar{\partial}}:\,\mathcal{W}(S)\to\mathcal{E}(S)

is unitary. Note that if μ(S)\mu\in\mathcal{E}(S), then by Green’s theorem μ\mu satisfies the neutrality condition (NC0):(\mathrm{NC}_{0}):

(3.4) μ=0.\int\mu=0.

Furthermore, μ2\|\mu\|_{\mathcal{E}}^{2} is represented as

μ2=×log1|zw|2μ(z)μ(w)¯\|\mu\|_{\mathcal{E}}^{2}=\iint_{\mathbb{C}\times\mathbb{C}}\log\frac{1}{|z-w|^{2}}\,\mu(z)\overline{\mu(w)}

for any ^\widehat{\mathbb{C}}-uniformization of S.S. The Aut(^)\mathrm{Aut}(\widehat{\mathbb{C}})-invariance of the right-hand side means that

×log1|τzτw|2μ(z)μ(w)¯\iint_{\mathbb{C}\times\mathbb{C}}\log\frac{1}{|\tau z-\tau w|^{2}}\,\mu(z)\overline{\mu(w)}

is independent of τAut(^).\tau\in\mathrm{Aut}(\widehat{\mathbb{C}}). Since translation invariance is trivial, it suffices to show this for τ(z)=az(a0)\tau(z)=az\,(a\neq 0) and τ(z)=1/z.\tau(z)=1/z. It follows from the neutrality condition (NC0)(\mathrm{NC}_{0}) on μ.\mu.

• Gaussian free field.  The Gaussian free field Ψ\Psi on SS is a (real, centered) Gaussian field indexed by (S)\mathcal{E}(S), i.e.,

Ψ:(S)L2(Ω,𝐏)\Psi:\,\mathcal{E}(S)\to L^{2}(\Omega,\mathbf{P})

is an isometry such that Ψ(μ)\Psi(\mu) is a (real, centered) Gaussian random variable for each μ(S).\mu\in\mathcal{E}(S). We introduce the Fock space functionals (“generalized” Gaussians)

Ψ(z,z0)=Ψ(δzδz0).\Psi(z,z_{0})=\Psi(\delta_{z}-\delta_{z_{0}}).

(Note that μ=δzδz0\mu=\delta_{z}-\delta_{z_{0}} is not in (S)\mathcal{E}(S) but as a signed/complex measure it can be approximated by μn\mu_{n}’s in \mathcal{E} because it satisfies (NC0);(\mathrm{NC}_{0}); here δ\delta’s are δ\delta-functions, i.e., (1,1)(1,1) forms.) Then the Fock space functionals Ψ(z,z0)\Psi(z,z_{0}) have the following properties:

  1. (a)

    Ψ(z,z0)=Ψ(z0,z),\Psi(z,z_{0})=-\Psi(z_{0},z), in particular Ψ(z,z)=0;\Psi(z,z)=0;

  2. (b)

    Ψ(z1,z3)=Ψ(z1,z2)+Ψ(z2,z3).\Psi(z_{1},z_{3})=\Psi(z_{1},z_{2})+\Psi(z_{2},z_{3}).

For μ(S),\mu\in\mathcal{E}(S), somewhat symbolically we have

Ψ(μ)=Ψ(z,z0)μ(z).\Psi(\mu)=\int\Psi(z,z_{0})\,\mu(z).

The integral is independent of the reference point z0z_{0} by (b) and the neutrality condition (NC0).(\mathrm{NC}_{0}). Thus we can think of the Gaussian free field as a bi-variant field with

𝐄Ψ(z,z0)Ψ(z,z0)=2log|(zz0)(z0z)(zz)(z0z0)|\mathbf{E}\,\Psi(z,z_{0})\,\Psi(z^{\prime},z_{0}^{\prime})=2\log\bigg{|}\frac{(z-z_{0}^{\prime})(z_{0}-z^{\prime})}{(z-z^{\prime})(z_{0}-z_{0}^{\prime})}\bigg{|}

on ^.\widehat{\mathbb{C}}. As a scalar, conformal invariance is manifest in the cross-ratio on the right-hand side. As in a simply-connected domain, we define the chiral boson Ψ+\Psi^{+} as a bi-variant multivalued field by

Ψ+(z,z0)=z0zζΨ(ζ,z1)dζ.\Psi^{+}(z,z_{0})=\int_{z_{0}}^{z}\partial_{\zeta}\Psi(\zeta,z_{1})\,\mathrm{d}\zeta.

It is obvious that Ψ+(z,z0)\Psi^{+}(z,z_{0}) does not depend on z1.z_{1}. On ^,\widehat{\mathbb{C}}, we have

𝐄Ψ+(z,z0)Ψ+(z,z0)=log(zz0)(z0z)(zz)(z0z0).\mathbf{E}\,\Psi^{+}(z,z_{0})\Psi^{+}(z^{\prime},z_{0}^{\prime})=\log\frac{(z-z_{0}^{\prime})(z_{0}-z^{\prime})}{(z-z^{\prime})(z_{0}-z_{0}^{\prime})}.

• Formal 1-point field.  We introduce Ψ\Psi on ^\widehat{\mathbb{C}} as a centered Gaussian formal field with formal correlation

𝐄Ψ(z1)Ψ(z2)=2log1|z1z2|.\mathbf{E}\,\Psi(z_{1})\,\Psi(z_{2})=2\log\frac{1}{|z_{1}-z_{2}|}.

For a compact Riemann surface SS of genus zero, we fix a uniformization w:S^w:S\to\widehat{\mathbb{C}} and define Ψ(z)=Ψ^(w(z))\Psi(z)=\Psi_{\widehat{\mathbb{C}}}(w(z)) so that

𝐄Ψ(z1)Ψ(z2)=2log1|w(z1)w(z2)|.\mathbf{E}\,\Psi(z_{1})\,\Psi(z_{2})=2\log\frac{1}{|w(z_{1})-w(z_{2})|}.

The dependence of the formal correlations on the choice of ww disappears if we apply this formalism only to the linear combinations satisfying the neutrality condition (NC0):(\mathrm{NC}_{0}):

jσjΨ(zj),σj=0.\sum_{j}\sigma_{j}\Psi(z_{j}),\qquad\sum\sigma_{j}=0.

For example, we have a representation Ψ(z,z0)=Ψ(z)Ψ(z0).\Psi(z,z_{0})=\Psi(z)-\Psi(z_{0}). Next we introduce the formal bosonic fields Ψ±Ψ^±\Psi^{\pm}\equiv\Psi^{\pm}_{\widehat{\mathbb{C}}} on ^\widehat{\mathbb{C}} as centered Gaussian formal fields satisfying Ψ=Ψ++Ψ,Ψ=Ψ+¯\Psi=\Psi^{+}+\Psi^{-},\Psi^{-}=\overline{\Psi^{+}} and

𝐄Ψ+(z1)Ψ+(z2)=log1z1z2,𝐄Ψ+(z1)Ψ(z2)=0.\mathbf{E}\,\Psi^{+}(z_{1})\,\Psi^{+}(z_{2})=\log\frac{1}{z_{1}-z_{2}},\qquad\mathbf{E}\,\Psi^{+}(z_{1})\,\Psi^{-}(z_{2})=0.

For a given uniformizing map w:S^,w:S\to\widehat{\mathbb{C}}, we define Ψ±(z)ΨS±(z):=Ψ^±(w(z)).\Psi^{\pm}(z)\equiv\Psi^{\pm}_{S}(z):=\Psi^{\pm}_{\widehat{\mathbb{C}}}(w(z)). If both 𝝈+=σj+zj\boldsymbol{\sigma}^{+}=\sum\sigma^{+}_{j}\cdot z_{j} and 𝝈=σjzj\boldsymbol{\sigma}^{-}=\sum\sigma^{-}_{j}\cdot z_{j} satisfy the neutrality condition (NC0),(\mathrm{NC}_{0}), then the linear combinations

Ψ+[𝝈+]:=σj+Ψ+(zj),Ψ[𝝈]:=σjΨ(zj),\Psi^{+}[\boldsymbol{\sigma}^{+}]:=\sum\sigma^{+}_{j}\Psi^{+}(z_{j}),\qquad\Psi^{-}[\boldsymbol{\sigma}^{-}]:=\sum\sigma^{-}_{j}\Psi^{-}(z_{j}),

and

Ψ[𝝈+,𝝈]:=Ψ+[𝝈+]Ψ[𝝈]=σj+Ψ+(zj)σjΨ(zj)\Psi[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]:=\Psi^{+}[\boldsymbol{\sigma}^{+}]-\Psi^{-}[\boldsymbol{\sigma}^{-}]=\sum\sigma^{+}_{j}\Psi^{+}(z_{j})-\sigma^{-}_{j}\Psi^{-}(z_{j})

are well-defined Fock space fields. For a divisor 𝝈=σjzj\boldsymbol{\sigma}^{-}=\sum\sigma^{-}_{j}\cdot z_{j} satisfying the neutrality condition (NC0),(\mathrm{NC}_{0}), we have

Ψ[𝝈]=Ψ+[𝝈¯]¯,\Psi^{-}[\boldsymbol{\sigma}^{-}]=\overline{\Psi^{+}[\overline{\boldsymbol{\sigma}^{-}}]},

where 𝝈¯=σj¯zj\overline{\boldsymbol{\sigma}}=\sum\overline{\sigma_{j}}\cdot z_{j} for 𝝈=σjzj.\boldsymbol{\sigma}=\sum\sigma_{j}\cdot z_{j}.

• Vertex fields.  Suppose that both 𝝈+\boldsymbol{\sigma}^{+} and 𝝈\boldsymbol{\sigma}^{-} satisfy the neutrality condition (NC0).(\mathrm{NC}_{0}). We define the multi-vertex field V[𝝈+,𝝈]V[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}] by

V[𝝈+,𝝈]=C[𝝈+,𝝈]V[𝝈+,𝝈],V[𝝈+,𝝈]:=eiΨ[𝝈+,𝝈],V[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]=C[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]V^{\odot}[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}],\qquad V^{\odot}[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]:=\mathrm{e}^{\odot i\Psi[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]},

where the Coulomb gas correlation function C[𝝈+,𝝈]CS[𝝈+,𝝈]C[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]\equiv C_{S}[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}] is given by

C[𝝈+,𝝈]=C[𝝈+]C[𝝈¯]¯.C[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]=C[\boldsymbol{\sigma}^{+}]\,\overline{C[\overline{\boldsymbol{\sigma}^{-}}]}.

We remark that 𝝈+\boldsymbol{\sigma}^{+} does not interact with 𝝈\boldsymbol{\sigma}^{-} in C[𝝈+,𝝈]C[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}] due to the independence of Ψ+\Psi^{+} and Ψ\Psi^{-}:

𝐄Ψ+(z1)Ψ(z2)=0.\mathbf{E}\,\Psi^{+}(z_{1})\,\Psi^{-}(z_{2})=0.

In the identity chart of \mathbb{C} and the chart z1/zz\mapsto-1/z at infinity, the value of the differential C[𝝈+,𝝈]C[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}] is

j<k(zjzk)σj+σk+(z¯jz¯k)σjσk,\prod_{j<k}(z_{j}-z_{k})^{\sigma_{j}^{+}\sigma_{k}^{+}}(\bar{z}_{j}-\bar{z}_{k})^{\sigma_{j}^{-}\sigma_{k}^{-}},

where the product is taken over finite zjz_{j}’s and zkz_{k}’s, and as usual 00:=1.0^{0}:=1.

Remark.

Recall that we may assume that σj=0\sigma_{j}^{-}=0 if zjD.z_{j}\in\partial D. There is a 1-1 correspondence between a divisor 𝝈\boldsymbol{\sigma} on S=DdoubleS=D^{\mathrm{double}} and a double divisor (𝝈+,𝝈)(\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}) with supp𝝈+DD,\mathrm{supp}\,\boldsymbol{\sigma}^{+}\subseteq D\cap\partial D, supp𝝈D:\mathrm{supp}\,\boldsymbol{\sigma}^{-}\subseteq D:

𝝈=𝝈++𝝈.\boldsymbol{\sigma}=\boldsymbol{\sigma}^{+}+\boldsymbol{\sigma}_{*}^{-}.

Sometimes it is convenient to write Φ[𝝈]\Phi[\boldsymbol{\sigma}] for Φ[𝝈+,𝝈],\Phi[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}], C[𝝈]C[\boldsymbol{\sigma}] for C[𝝈+,𝝈],C[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}], and V[𝝈]V[\boldsymbol{\sigma}] for V[𝝈+,𝝈],V[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}], etc.

4. Modified multi-vertex fields

In this section we extend the concept of multi-vertex fields to the case b0.b\neq 0. Their correlation functions are defined in terms of correlation functions with the neutrality condition (NCb)(\mathrm{NC}_{b}) and their Wick’s parts are Wick’s exponentials of bosonic fields with charges 𝝉\boldsymbol{\tau} satisfying the neutrality condition (NC0).(\mathrm{NC}_{0}). To reconcile these two neutrality conditions, background charges 𝜷\boldsymbol{\beta} with the neutrality condition (NCb)(\mathrm{NC}_{b}) are placed. In the next section we view the modified multi-vertex fields 𝒪𝜷[𝝉]\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}] as the OPE exponentials of background charge modification of iΦ[𝝉].i\Phi[\boldsymbol{\tau}].

4.1. Definition

Let us fix the (background charge) parameter b.b\in\mathbb{R}. This parameter bb is related to the central charge cc in the following way:

c=112b2.c=1-12b^{2}.

As we mentioned in the last remark in the previous section, we often use the 1-1 correspondence between a divisor 𝝈\boldsymbol{\sigma} on S=DdoubleS=D^{\mathrm{double}} and a double divisor (𝝈+,𝝈)(\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}) with supp𝝈+DD,\mathrm{supp}\,\boldsymbol{\sigma}^{+}\subseteq D\cap\partial D, supp𝝈D:\mathrm{supp}\,\boldsymbol{\sigma}^{-}\subseteq D: 𝝈=𝝈++𝝈.\boldsymbol{\sigma}=\boldsymbol{\sigma}^{+}+\boldsymbol{\sigma}_{*}^{-}. Recall that the Coulomb gas correlation functions C[𝝈]C(b)[𝝈+,𝝈]C[\boldsymbol{\sigma}]\equiv C_{(b)}[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}] are well-defined [𝝀+,𝝀][\boldsymbol{\lambda}^{+},\boldsymbol{\lambda}^{-}]-differentials under the neutrality condition (NCb):(\mathrm{NC}_{b}):

𝝈=2b,\int\boldsymbol{\sigma}=2b,

see Theorem 2.3, and that Wick’s exponentials V[𝝉]V[𝝉+,𝝉]=eiΦ[𝝉+,𝝉]V^{\odot}[\boldsymbol{\tau}]\equiv V^{\odot}[\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}]=\mathrm{e}^{\odot i\Phi[\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}]} (𝝉=𝝉++𝝉)(\boldsymbol{\tau}=\boldsymbol{\tau}^{+}+\boldsymbol{\tau}_{*}^{-}) are well-defined Fock space fields under the neutrality condition (NC0):(\mathrm{NC}_{0}):

𝝉=0,\int\boldsymbol{\tau}=0,

where Φ[𝝉]Φ[𝝉+,𝝉]=τj+Φ+(zj)τjΦ(zj),\Phi[\boldsymbol{\tau}]\equiv\Phi[\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}]=\sum\tau_{j}^{+}\Phi^{+}(z_{j})-\tau_{j}^{-}\Phi^{-}(z_{j}), see Subsection 3.4.

To reconcile these two neutrality conditions (NCb)(\mathrm{NC}_{b}) on C(b)[𝝈+,𝝈]C_{(b)}[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}] and (NC0)(\mathrm{NC}_{0}) on V[𝝉+,𝝉],V^{\odot}[\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}], we need at least one marked point in DDD\cup\partial D to place the “background charge” there. Let us consider the case of one marked point and denote by qq this marked point. It can be one of the nodes of (𝝈+,𝝈).(\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}). The case qDq\in D is called radial, and the case qDq\in\partial D chordal.

• Standard chordal case.  For a divisor 𝝈(=𝝈++𝝈)\boldsymbol{\sigma}(=\boldsymbol{\sigma}^{+}+\boldsymbol{\sigma}_{*}^{-}) on SS satisfying the neutrality condition (NCb)(\mathrm{NC}_{b}) we define V[𝝈]V[𝝈+,𝝈]V[\boldsymbol{\sigma}]\equiv V[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}] by

V[𝝈]=C(b)[𝝈]V[𝝈2bq].V[\boldsymbol{\sigma}]=C_{(b)}[\boldsymbol{\sigma}]~{}V^{\odot}[\boldsymbol{\sigma}-2b\cdot q].

Sometimes we write VqV_{q} or V(b,q)V_{(b,q)} to indicate the location of the marked point and/or the background charge. Let us emphasize that the expectation of the multi-vertex (the correlation C(b)[𝝈]C_{(b)}[\boldsymbol{\sigma}]) does not depend on qq but the Wick part does. Multi-vertex functionals/fields are differentials with the same conformal dimensions as in C(b)[𝝈].C_{(b)}[\boldsymbol{\sigma}].

• Standard radial case.  For a divisor 𝝈(=𝝈++𝝈)\boldsymbol{\sigma}(=\boldsymbol{\sigma}^{+}+\boldsymbol{\sigma}_{*}^{-}) on SS satisfying the neutrality condition (NCb)(\mathrm{NC}_{b}) we define V[𝝈]V[𝝈+,𝝈]V[\boldsymbol{\sigma}]\equiv V[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}] by

V[𝝈]=C(b)[𝝈]V[𝝈bqbq].V[\boldsymbol{\sigma}]=C_{(b)}[\boldsymbol{\sigma}]~{}V^{\odot}[\boldsymbol{\sigma}-b\cdot q-b\cdot q^{*}].

Again, V[𝝈]V[\boldsymbol{\sigma}] is Aut(D,q)\mathrm{Aut}(D,q)-invariant but its expectation C(b)[𝝈]C_{(b)}[\boldsymbol{\sigma}] is Aut(D)\mathrm{Aut}(D)-invariant. (This is not the only way to satisfy neutrality condition in Wick’s part of the functional.)

Both cases can be generalized to several marked points. Additional marked points qkq_{k} are not necessarily on the boundary D.\partial D. Suppose we have a background charge 𝜷(=𝜷++𝜷)\boldsymbol{\beta}(=\boldsymbol{\beta}^{+}+\boldsymbol{\beta}_{*}^{-}) on SS with the neutrality condition (NCb).(\mathrm{NC}_{b}). For a divisor 𝝈(=𝝈++𝝈)\boldsymbol{\sigma}(=\boldsymbol{\sigma}^{+}+\boldsymbol{\sigma}_{*}^{-}) on SS satisfying the neutrality condition (NCb)(\mathrm{NC}_{b}) we define V𝜷[𝝈]V𝜷+,𝜷[𝝈+,𝝈]V_{\boldsymbol{\beta}}[\boldsymbol{\sigma}]\equiv V_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}] by

V𝜷[𝝈]=C(b)[𝝈]V[𝝈𝜷].V_{\boldsymbol{\beta}}[\boldsymbol{\sigma}]=C_{(b)}[\boldsymbol{\sigma}]~{}V^{\odot}[\boldsymbol{\sigma}-\boldsymbol{\beta}].

4.2. Background charge operators

Let us introduce the background charge operators. For a background charge 𝜷\boldsymbol{\beta} on S,S, we define the background charge operator 𝒫𝜷\mathcal{P}_{\boldsymbol{\beta}} associated with 𝜷\boldsymbol{\beta} by

𝒫𝜷:=V𝜷[𝜷]=C(b)[𝜷].\mathcal{P}_{\boldsymbol{\beta}}:=V_{\boldsymbol{\beta}}[\boldsymbol{\beta}]=C_{(b)}[\boldsymbol{\beta}].

Note that the Wick part of V[𝜷]V[\boldsymbol{\beta}] is the constant field 1.1.

• Standard chordal case.  We write 𝒫q\mathcal{P}_{q} for 𝒫2bq=C(b)[2bq].\mathcal{P}_{2b\cdot q}=C_{(b)}[2b\cdot q]. Then we would have 𝒫q=1\mathcal{P}_{q}=1 in all charts because

λq+=λb(2b)=0\lambda_{q}^{+}=\lambda_{b}(2b)=0

and clearly 𝒫q=1\mathcal{P}_{q}=1 in the (,)(\mathbb{H},\infty)-uniformization.

• Standard radial case.  We write 𝒫q\mathcal{P}_{q} for 𝒫bq+bq=C(b)[bq+bq].\mathcal{P}_{b\cdot q+b\cdot q^{*}}=C_{(b)}[b\cdot q+b\cdot q^{*}]. This is a non-random differential with dimensions

λq+=λq=λb(b)=b22\lambda_{q}^{+}=\lambda_{q}^{-}=\lambda_{b}(b)=-\frac{b^{2}}{2}

at q.q. Note that 𝒫q=1\mathcal{P}_{q}=1 in the (𝔻,0)(\mathbb{D},0)-uniformization.

4.3. OPE exponentials

We use the background charge operators to modify the multi-vertex fields so that the OPE calculus of the modified multi-vertex fields has a very simple and natural form. We call the modified multi-vertex fields the OPE exponentials; the reason for this terminology becomes apparent in the next section.

Suppose 𝝉(=𝝉++𝝉)\boldsymbol{\tau}(=\boldsymbol{\tau}^{+}+\boldsymbol{\tau}_{*}^{-}) is a divisor on SS satisfying the neutrality condition (NC0).(\mathrm{NC}_{0}). For this divisor 𝝉\boldsymbol{\tau} and a background charge 𝜷(=𝜷++𝜷)\boldsymbol{\beta}(=\boldsymbol{\beta}^{+}+\boldsymbol{\beta}_{*}^{-}) on SS (satisfying the neutrality condition (NCb)(\mathrm{NC}_{b})), we define the OPE exponentials (the modified multi-vertex fields) 𝒪𝜷[𝝉]𝒪𝜷+,𝜷[𝝉+,𝝉]\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}]\equiv\mathcal{O}_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}[\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}] by

𝒪𝜷[𝝉]:=𝒫𝜷1V𝜷[𝝉+𝜷]=C(b)[𝝉+𝜷]C(b)[𝜷]V[𝝉].\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}]:=\mathcal{P}_{\boldsymbol{\beta}}^{-1}V_{\boldsymbol{\beta}}[\boldsymbol{\tau}+\boldsymbol{\beta}]=\frac{C_{(b)}[\boldsymbol{\tau}+\boldsymbol{\beta}]}{C_{(b)}[\boldsymbol{\beta}]}~{}V^{\odot}[\boldsymbol{\tau}].

We remark that all charges in the correlation

𝐄𝒪𝜷[𝝉]=C(b)[𝝉+𝜷]C(b)[𝜷]\mathbf{E}\,\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}]=\frac{C_{(b)}[\boldsymbol{\tau}+\boldsymbol{\beta}]}{C_{(b)}[\boldsymbol{\beta}]}

interact except that background charges do not interact with each other.

• Standard radial case.  If 𝜷=bq+bq,\boldsymbol{\beta}=b\cdot q+b\cdot q^{*}, then

𝒪[𝝉]𝒪𝜷[𝝉]=C(b)[𝝉+bq+bq]C(b)[bq+bq]V[𝝉].\mathcal{O}[\boldsymbol{\tau}]\equiv\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}]=\frac{C_{(b)}[\boldsymbol{\tau}+b\cdot q+b\cdot q^{*}]}{C_{(b)}[b\cdot q+b\cdot q^{*}]}~{}V^{\odot}[\boldsymbol{\tau}].

We call (𝝉+,𝝉)(\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}) the (double) divisor of Wick’s exponents and (τq++b,τq+b)(\tau_{q}^{+}+b,\tau_{q}^{-}+b) the effective charges at q.q. Thus the sum of exponents is zero but the sum of (effective) charges is 2b,2b, i.e., 𝝉\boldsymbol{\tau} satisfies (NC0)(\mathrm{NC}_{0}) and 𝝉+𝜷\boldsymbol{\tau}+\boldsymbol{\beta} satisfies (NCb).(\mathrm{NC}_{b}). Sometimes we omit the subscript 𝜷\boldsymbol{\beta} in 𝒪𝜷[𝝉]\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}] when 𝜷=bq+bq.\boldsymbol{\beta}=b\cdot q+b\cdot q^{*}.

Proposition 4.1.

The conformal dimensions of 𝒪[𝛕]\,\mathcal{O}[\boldsymbol{\tau}] at qq are

hq±=(τq±)22.h_{q}^{\pm}=\frac{(\tau_{q}^{\pm})^{2}}{2}.
Proof.

The Coulomb gas correlations C(b)[𝝉++bq,𝝉+bq],C(b)[bq,bq]C_{(b)}[\boldsymbol{\tau}^{+}+b\cdot q,\boldsymbol{\tau}^{-}+b\cdot q],C_{(b)}[b\cdot q,b\cdot q] have conformal dimensions [λb(τq++b),λb(τq+b)],[\lambda_{b}(\tau_{q}^{+}+b),\lambda_{b}(\tau_{q}^{-}+b)], [λb(b),λb(b)],[\lambda_{b}(b),\lambda_{b}(b)], respectively at q.q. Since Wick’s exponentials V[𝝉+,𝝉]V^{\odot}[\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}] are scalars or (0,0)(0,0)-differentials, we have

hq±=λb(τq±+b)λb(b)=(τq±)22.h_{q}^{\pm}=\lambda_{b}(\tau_{q}^{\pm}+b)-\lambda_{b}(b)=\frac{(\tau_{q}^{\pm})^{2}}{2}.

We call the numbers λb(τq±+b)\lambda_{b}(\tau_{q}^{\pm}+b) effective conformal dimensions at qq; they are dimensions of 𝒫q𝒪[𝝉]=V[𝝉+𝜷].\mathcal{P}_{q}\,\mathcal{O}[\boldsymbol{\tau}]=V[\boldsymbol{\tau}+\boldsymbol{\beta}]. As we have just seen,

λb(τq±+b)=hq±b22.\lambda_{b}(\tau_{q}^{\pm}+b)=h_{q}^{\pm}-\frac{b^{2}}{2}.
Proposition 4.2.

In the (𝔻,0)(\mathbb{D},0)-uniformization, we have

𝐄𝒪[𝝉]=j(zj)νj+(z¯j)νjj<k(zjzk)τj+τk+(z¯jz¯k)τjτkj,k(1zjz¯k)τj+τk,\mathbf{E}\,\mathcal{O}[\boldsymbol{\tau}]=\prod_{j}(z_{j})^{\nu_{j}^{+}}(\bar{z}_{j})^{\nu_{j}^{-}}\prod_{j<k}(z_{j}-z_{k})^{\tau_{j}^{+}\tau_{k}^{+}}(\bar{z}_{j}-\bar{z}_{k})^{\tau_{j}^{-}\tau_{k}^{-}}\prod_{j,k}(1-z_{j}\bar{z}_{k})^{\tau_{j}^{+}\tau_{k}^{-}},

where νj±=τj±(τq±+b)\nu_{j}^{\pm}=\tau_{j}^{\pm}(\tau_{q}^{\pm}+b) and zjz_{j}’s are non-zero nodes.

Proof.

Since 𝒫q=1\mathcal{P}_{q}=1 in the (𝔻,0)(\mathbb{D},0)-uniformization and 𝐄V[𝝉]=1,\mathbf{E}\,V^{\odot}[\boldsymbol{\tau}]=1, we have

𝐄𝒪[𝝉]=C(b)[𝝉+bq+bq].\mathbf{E}\,\mathcal{O}[\boldsymbol{\tau}]=C_{(b)}[\boldsymbol{\tau}+b\cdot q+b\cdot q^{*}].

Proposition now follows from Theorem 2.5. We have 𝐄𝒪[𝝉]=C𝔻[𝝉++bq,𝝉+bq]\mathbf{E}\,\mathcal{O}[\boldsymbol{\tau}]=C_{\mathbb{D}}[\boldsymbol{\tau}^{+}+b\cdot q,\boldsymbol{\tau}^{-}+b\cdot q] in the identity chart of 𝔻.\mathbb{D}.

In other words, 𝐄𝒪[𝝉]\mathbf{E}\,\mathcal{O}[\boldsymbol{\tau}] in (𝔻,0)(\mathbb{D},0) is given by the usual Coulomb gas correlation function for effective charges.

• Standard chordal case.  If 𝜷=2bq,\boldsymbol{\beta}=2b\cdot q, then

𝒪[𝝉]𝒪𝜷[𝝉]:=V[𝝉+2bq]=C(b)[𝝉+2bq]V[𝝉].\mathcal{O}[\boldsymbol{\tau}]\equiv\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}]:=V[\boldsymbol{\tau}+2b\cdot q]=C_{(b)}[\boldsymbol{\tau}+2b\cdot q]\,V^{\odot}[\boldsymbol{\tau}].

Recall that 𝒫q=1\mathcal{P}_{q}=1 in all charts. The (effective) charge at qq is τq++2b,\tau_{q}^{+}+2b, so the sum of (effective) charges is 2b.2b. Sometimes we omit subscript 𝜷\boldsymbol{\beta} in 𝒪𝜷[𝝉]\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}] when 𝜷=2bq.\boldsymbol{\beta}=2b\cdot q. The conformal dimension of 𝒪[𝝉]\mathcal{O}[\boldsymbol{\tau}] at qq is

hq+=λb(τq++2b)=λb(τq+)=(τq+)22+τq+b.h_{q}^{+}=\lambda_{b}(\tau_{q}^{+}+2b)=\lambda_{b}(-\tau_{q}^{+})=\frac{(\tau_{q}^{+})^{2}}{2}+\tau_{q}^{+}b.
Remark.

In the standard chordal case, there is a 1-to-1 correspondence between the double divisors (𝝈+,𝝈)(\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}) in D¯{q}\mkern 1.5mu\overline{\mkern-1.5muD\mkern-1.5mu}\mkern 1.5mu\setminus\{q\} and the double divisors (𝝉+,𝝉)(\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}) in D¯\mkern 1.5mu\overline{\mkern-1.5muD\mkern-1.5mu}\mkern 1.5mu satisfying (NC0)(\mathrm{NC}_{0}) with τq=0:\tau_{q}^{-}=0:

𝝈+=𝝉+τq+q,𝝈=𝝉,τq+=(𝝈++𝝈).\boldsymbol{\sigma}^{+}=\boldsymbol{\tau}^{+}-\tau_{q}^{+}\cdot q,\quad\boldsymbol{\sigma}^{-}=\boldsymbol{\tau}^{-},\quad\tau_{q}^{+}=-\int(\boldsymbol{\sigma}^{+}+\boldsymbol{\sigma}^{-}).

In [18] we use the notation

Σ=(𝝈++𝝈),\Sigma=\int(\boldsymbol{\sigma}^{+}+\boldsymbol{\sigma}^{-}),

so Σ=τq+\Sigma=-\tau_{q}^{+} (by (NC0)(\mathrm{NC}_{0})) and hq=λb(Σ).h_{q}=\lambda_{b}(\Sigma). Also, in [18] we use the notation

𝒪(σ1+,σ1)(z1)𝒪(σn+,σn)(zn)\mathcal{O}^{(\sigma_{1}^{+},\sigma_{1}^{-})}(z_{1})\star\cdots\star\mathcal{O}^{(\sigma_{n}^{+},\sigma_{n}^{-})}(z_{n})

(which can be shortened to 𝒪(𝝈+,𝝈)\mathcal{O}^{(\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-})}) instead of our present notation 𝒪[𝝉+,𝝉]\mathcal{O}[\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}] for the OPE exponentials.

4.4. Example: one-leg operators

As a special case of Theorem 1.5, under the insertion of Wick’s part Λp/𝐄Λp\Lambda_{p}/\mathbf{E}\,\Lambda_{p} of the radial one-leg operator Λp\Lambda_{p} (rooted at qq) with pDp\in\partial D all correlation functions of the fields in the extended OPE family of Φ𝜷\Phi_{\boldsymbol{\beta}} (𝜷=bq+bq\boldsymbol{\beta}=b\cdot q+b\cdot q^{*}) are radial SLE(κ)\mathrm{SLE}(\kappa) martingale-observables (see Subsection 5.4 for the modified Gaussian free Φ𝜷\Phi_{\boldsymbol{\beta}} and Subsection 6.2 for its extended OPE family) if the parameters aa (the charge of Λp\Lambda_{p} at pp) and bb are related to the SLE parameter κ\kappa as

a=±2/κ,b=a(κ/41).a=\pm\sqrt{2/\kappa},\qquad b=a(\kappa/4-1).

(a) We define the radial one-leg operator Λ\Lambda (rooted at qq) by

ΛzΛ(z):=𝒪[aza2qa2q].\Lambda_{z}\equiv\Lambda(z):=\mathcal{O}\big{[}a\cdot z-\frac{a}{2}\cdot q-\frac{a}{2}\cdot q^{*}\big{]}.

Its conformal dimensions λz\lambda_{z} at zz and hq±h_{q}^{\pm} at qq are

(4.1) λz=hh1,2:=a22ab=6κ2κ\lambda_{z}=h\equiv h_{1,2}:=\frac{a^{2}}{2}-ab=\frac{6-\kappa}{2\kappa}

and

hq+=hq=a28,Hq(:=hq++hq)=a24.h_{q}^{+}=h_{q}^{-}=\frac{a^{2}}{8},\qquad H_{q}(:=h_{q}^{+}+h_{q}^{-})=\frac{a^{2}}{4}.

The effective charges at qq are (ba/2,ba/2)(b-a/2,b-a/2) and the effective dimensions at qq are

(4.2) Hqeff(Λ)=2h0,1/2,h0,1/2:=a28b22=(κ2)(6κ)16κ.H_{q}^{\mathrm{eff}}(\Lambda)=2h_{0,1/2},\quad h_{0,1/2}:=\frac{a^{2}}{8}-\frac{b^{2}}{2}=\frac{(\kappa-2)(6-\kappa)}{16\kappa}.

Indeed, Hqeff(Λ)=Hq(Λeff),H_{q}^{\mathrm{eff}}(\Lambda)=H_{q}(\Lambda^{\mathrm{eff}}), where Λeff\Lambda^{\mathrm{eff}} is the “effective” one-leg operator

Λeff=𝒫qΛ=V[az+(ba2)q+(ba2)q].\Lambda^{\mathrm{eff}}=\mathcal{P}_{q}\,\Lambda=V\big{[}a\cdot z+\big{(}b-\frac{a}{2}\big{)}\cdot q+\big{(}b-\frac{a}{2}\big{)}\cdot q^{*}\big{]}.

In (𝔻,0)(\mathbb{D},0) we have 𝐄Λ(z)=𝐄Λeff(z)=za(ba/2)=zh,\mathbf{E}\,\Lambda(z)=\mathbf{E}\,\Lambda^{\mathrm{eff}}(z)=z^{a(b-a/2)}=z^{-h}, so in (D,q)(D,q)

𝐄Λ(z)=(w(z)w(z))h|wq|Hq,\mathbf{E}\,\Lambda(z)=\Big{(}\frac{w^{\prime}(z)}{w(z)}\Big{)}^{h}|w_{q}^{\prime}|^{H_{q}},

where w:(D,q)(𝔻,0)w:(D,q)\to(\mathbb{D},0) is a conformal map and wq=w(q).w_{q}^{\prime}=w^{\prime}(q).

(b) Let η\eta\in\mathbb{R} be the parameter in the definition of radial SLEη(κ,𝝆),\mathrm{SLE}_{\eta}(\kappa,\boldsymbol{\rho}), see (1.8). We define the radial one-leg operator Λ(s)\Lambda_{(s)} with spin s=iηa2/2s=i\eta a^{2}/2 by

Λ(s)(z):=𝒪[aza+iδ2qaiδ2q],\Lambda_{(s)}(z):=\mathcal{O}\big{[}a\cdot z-\frac{a+i\delta}{2}\cdot q-\frac{a-i\delta}{2}\cdot q^{*}\big{]},

where δ=ηa.\delta=\eta a. Its conformal dimensions hq±h_{q}^{\pm} at qq are

hq±=(a±iδ)28,h_{q}^{\pm}=\frac{(a\pm i\delta)^{2}}{8},

so the spin s:=hq+hqs:=h_{q}^{+}-h_{q}^{-} and the conformal dimension Hq:=hq++hqH_{q}:=h_{q}^{+}+h_{q}^{-} at qq are

s=iaδ2,Hq=a2+δ24.s=\frac{ia\delta}{2},\qquad H_{q}=\frac{a^{2}+\delta^{2}}{4}.

In (𝔻,0)(\mathbb{D},0) we have 𝐄Λ(s)(z)=𝐄Λ(s)eff(z)=za(ba/2iδ/2)=zhs,\mathbf{E}\,\Lambda_{(s)}(z)=\mathbf{E}\,\Lambda_{(s)}^{\mathrm{eff}}(z)=z^{a(b-a/2-i\delta/2)}=z^{-h-s}, and

κθlog𝐄Λ(s)eff(eiθ)𝐄Λeff(eiθ)=η.\kappa\partial_{\theta}\log\frac{\mathbf{E}\,\Lambda_{(s)}^{\mathrm{eff}}(\mathrm{e}^{i\theta})}{\mathbf{E}\,\Lambda^{\mathrm{eff}}(\mathrm{e}^{i\theta})}=\eta.

See (1.7) and (1.8). Later we define the partition function in terms of the correlation function of the effective one-leg operator.

(c) We define the chordal one-leg operator Λ\Lambda by

Λ(z):=𝒪[azaq].\Lambda(z):=\mathcal{O}\big{[}a\cdot z-a\cdot q].

Its conformal dimension λz\lambda_{z} at zz and effective dimension hqeffh_{q}^{\mathrm{eff}} at qq are

λz=h,hqeff=λb(2ba)=λb(a)=h.\lambda_{z}=h,\qquad h_{q}^{\mathrm{eff}}=\lambda_{b}(2b-a)=\lambda_{b}(a)=h.

In (,)(\mathbb{H},\infty) we have 𝐄Λeff(z)=1,\mathbf{E}\,\Lambda^{\mathrm{eff}}(z)=1, so 𝐄Λeff(z)=(w)h(wq)h\mathbf{E}\,\Lambda^{\mathrm{eff}}(z)=(w^{\prime})^{h}(w_{q}^{\prime})^{h} in (D,q).(D,q).

4.5. Algebra of multi-vertex functionals and OPE exponentials

In the general case, given a background charge 𝜷\boldsymbol{\beta} we define the multiplication of OPE exponentials by

𝒪𝜷[𝝉1]𝒪𝜷[𝝉2]=𝒪𝜷[𝝉1+𝝉2].\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}_{1}]\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}_{2}]=\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}_{1}+\boldsymbol{\tau}_{2}].

In the next section we relate this operation to OPE/tensor multiplication. Multiplication of OPE exponentials is commutative and associative (if we ignore the order of particles). Formally, we have the following representation

𝒪𝜷[𝝉+,𝝉]=𝒪𝜷(τ1+)(z1)𝒪𝜷(τ¯1)(z1)¯𝒪𝜷(τn+)(zn)𝒪𝜷(τ¯n)(zn)¯,\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}]=\mathcal{O}_{\boldsymbol{\beta}}^{(\tau_{1}^{+})}(z_{1})\overline{\mathcal{O}_{\boldsymbol{\beta}}^{(\bar{\tau}_{1}^{-})}(z_{1})}\cdots\mathcal{O}_{\boldsymbol{\beta}}^{(\tau_{n}^{+})}(z_{n})\overline{\mathcal{O}_{\boldsymbol{\beta}}^{(\bar{\tau}_{n}^{-})}(z_{n})},

where τ¯j=τj¯.\bar{\tau}_{j}^{-}=\overline{\tau_{j}^{-}}.

5. Background charge modifications of Gaussian free field

In this section we discuss background charge modifications of the Gaussian free field in a simply-connected domain DD with marked boundary/interior points. A special type of such modifications with a marked boundary point appeared in [29] and [18] to present their connections to chordal SLE theory. Similar constructions had been well known both in the physics literature and in the algebraic literature. For example, see [11, Chapter 9] for Coulomb gas formalism, and [16] for Fairlie’s modifications of Virasoro generators.

The background charge modifications of the radial conformal field theory with a marked interior point qq under our consideration equip the Gaussian free field with the additive monodromy around the puncture qD.q\in D. We define the OPE exponentials of modified bosonic fields with nodes in D=D{q}D^{*}=D\setminus\{q\} as differentials in the OPE family. The OPE exponentials extend to the puncture through operator product expansion with the constant field 1q1_{q} or the rooting procedure. We discuss these extensions in the next section.

5.1. Modification of Gaussian free field on the Riemann sphere

In this subsection we borrow the concepts of the background charge modifications of Gaussian free field on a compact Riemann surface from [17]. Later we adapt them in a simply-connected domain employing Schottky double construction.

Recall that a non-random field ψ\psi is called a pre-pre-Schwarzian form of order (μ,ν)(\mu,\nu) (or PPS(μ,ν)\mathrm{PPS}(\mu,\nu)) if the transformation law is

ψ=ψ~h+μlogh+νlogh¯,\psi=\tilde{\psi}\circ h+\mu\log h^{\prime}+\nu\log\overline{h^{\prime}},

where ψ=(ψϕ)\psi=(\psi\,\|\,\phi) in a chart ϕ,\phi, ψ~=(ψϕ~)\tilde{\psi}=(\psi\,\|\,\tilde{\phi}) in a chart ϕ~,\tilde{\phi}, and hh is the transition map between two overlapping charts ϕ,ϕ~.\phi,\tilde{\phi}. We consider a holomorphic/harmonic PPS form ψ\psi such that ψ\partial\psi is meromorphic and 𝜷𝜷ψ:=i/π¯ψ\boldsymbol{\beta}\equiv\boldsymbol{\beta}_{\psi}:=i/\pi{\bar{\partial}}\partial\psi is a finite linear combination of δ\delta-measures, 𝜷=kβkδqk.\boldsymbol{\beta}=\sum_{k}\beta_{k}\delta_{q_{k}}. Such ψ\psi is called a simple PPS form. We often write 𝜷ψ=kβkqk\boldsymbol{\beta}_{\psi}=\sum_{k}\beta_{k}\cdot q_{k} as a divisor and call it the background charge of ψ.\psi.

Proposition 5.1.

On a compact Riemann surface SS of genus zero, we have the neutrality condition (NCb)(\mathrm{NC}_{b}) for the background charge 𝛃ψ\boldsymbol{\beta}_{\psi} of a simple PPS(ib,0)\mathrm{PPS}(ib,0) form ψ:\psi:

(5.1) kβk=2b.\sum_{k}\beta_{k}=2b.

It is a consequence of a version of Gauss-Bonnet theorem, e.g., see [17, Corollary 6.3]. For the reader’s convenience, we present its proof.

Proof of Proposition 5.1.

Let us choose a conformal metric ρ=|ωq|2\rho=|\omega_{q}|^{2} (a positive (1,1)(1,1)-differential) on S,S, where ωq\omega_{q} is a meromorphic differential with a sole double pole at qS.q\in S. In the identity chart of \mathbb{C} with q=0,q=0, one can take ρ(z)=|z|4,\rho(z)=|z|^{-4}, so ¯logρ=2πδ0{\bar{\partial}}\partial\log\rho=-2\pi\delta_{0} and

S¯logρ=2π=πχ(S),\int_{S}{\bar{\partial}}\partial\log\rho=-2\pi=-\pi\chi(S),

where χ(S)\chi(S) is the Euler characteristic of S.S. Let φ=ψ𝜷/(ib).\varphi=\psi_{\boldsymbol{\beta}}/(ib). Then φ\varphi is a simple PPS(1,0)\mathrm{PPS}(1,0) form. We now consider the harmonic PPS(1,1)\mathrm{PPS}(1,1) form, φ=logρ=log|ω0|2.\varphi_{*}=\log\rho=\log|\omega_{0}|^{2}. Proposition now follows from

(5.2) S¯(φφ)=0\int_{S}{\bar{\partial}}\partial(\varphi-\varphi_{*})=0

since

𝜷=iπ¯ψ𝜷=iπib¯φ=iπib¯φ=2b.\int\boldsymbol{\beta}=\frac{i}{\pi}\int{\bar{\partial}}\partial\psi_{\boldsymbol{\beta}}=\frac{i}{\pi}ib\int{\bar{\partial}}\partial\varphi=\frac{i}{\pi}ib\int{\bar{\partial}}\partial\varphi_{*}=2b.

By means of Green’s theorem, the integral in (5.2) is the sum of all residues of the meromorphic differential (φφ)/(2i).\partial(\varphi-\varphi_{*})/(2i). It is well known that such a sum vanishes. ∎

Given a background charge 𝜷=kβkqk\boldsymbol{\beta}=\sum_{k}\beta_{k}\cdot q_{k} with the neutrality condition (NCb)(\mathrm{NC}_{b}), there is a unique (up to an additive constant) simple PPS(ib,0)\mathrm{PPS}(ib,0) form ψ𝜷+\psi_{\boldsymbol{\beta}}^{+} with the background charge 𝜷,\boldsymbol{\beta},

ψ𝜷+=ikβkψqk+,\psi_{\boldsymbol{\beta}}^{+}=i\sum_{k}\beta_{k}\psi_{q_{k}}^{+},

where ψq+=12logωq\psi_{q}^{+}=\frac{1}{2}\log\omega_{q} and ωq\omega_{q} is a meromorphic differential with a sole double pole at qS.q\in S. On the Riemann sphere ^,\widehat{\mathbb{C}}, ω=1\omega_{\infty}=1 and ωq=1/(zq)2\omega_{q}=1/(z-q)^{2} in the identity chart of .\mathbb{C}. In terms of a uniformization w:S^,w:S\to\widehat{\mathbb{C}},

ψqk+(z)={12logw(z)log(w(z)w(qk)),w(qk)12logw(z),w(qk)=.\psi_{q_{k}}^{+}(z)=\begin{cases}\frac{1}{2}\log w^{\prime}(z)-\log(w(z)-w(q_{k})),\qquad&w(q_{k})\neq\infty\\ \frac{1}{2}\log w^{\prime}(z),&w(q_{k})=\infty.\end{cases}

We now define the background charge modifications of the formal bosonic fields. Given two background charges 𝜷±=kβk±qk\boldsymbol{\beta}^{\pm}=\sum_{k}\beta_{k}^{\pm}\cdot q_{k} with the neutrality conditions βk±=2b±,\sum\beta_{k}^{\pm}=2b^{\pm}, we define

Ψ𝜷±±=Ψ±+ψ𝜷±±,ψ𝜷=ψ𝜷¯+¯,Ψ𝜷+,𝜷=Ψ𝜷+++Ψ𝜷,ψ𝜷+,𝜷=ψ𝜷+++ψ𝜷\Psi_{\boldsymbol{\beta}^{\pm}}^{\pm}=\Psi^{\pm}+\psi_{\boldsymbol{\beta}^{\pm}}^{\pm},\qquad\psi_{\boldsymbol{\beta}^{-}}^{-}=\overline{\psi_{\mkern 1.5mu\overline{\mkern-1.5mu\boldsymbol{\beta}^{-}\mkern-1.5mu}\mkern 1.5mu}^{+}},\qquad\Psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}=\Psi_{\boldsymbol{\beta}^{+}}^{+}+\Psi_{\boldsymbol{\beta}^{-}}^{-},\qquad\psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}=\psi_{\boldsymbol{\beta}^{+}}^{+}+\psi_{\boldsymbol{\beta}^{-}}^{-}

so that Ψ𝜷+,𝜷=Ψ+ψ𝜷+,𝜷.\Psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}=\Psi+\psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}. Then Ψ𝜷+,𝜷\Psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}} is a PPS(ib+,ib)\mathrm{PPS}(ib^{+},-ib^{-}) form. In the \mathbb{C}-uniformization, we have

ψ𝜷+,𝜷(z)\displaystyle\psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}(z) =ikβk+log1zqkikβklog1z¯q¯k\displaystyle=i\sum_{k}\beta_{k}^{+}\log\frac{1}{z-q_{k}}-i\sum_{k}\beta_{k}^{-}\log\frac{1}{\bar{z}-\bar{q}_{k}}
=ik(βk+βk)log1|zqk|+k(βk++βk)arg(zqk).\displaystyle=i\sum_{k}(\beta_{k}^{+}-\beta_{k}^{-})\log\frac{1}{|z-q_{k}|}+\sum_{k}(\beta_{k}^{+}+\beta_{k}^{-})\arg(z-q_{k}).

Here the summation is taken over all kk’s but qk=.q_{k}=\infty. In particular, if 𝜷+=𝜷¯𝜷,\boldsymbol{\beta}^{+}=\overline{\boldsymbol{\beta}^{-}}\equiv\boldsymbol{\beta}, then

(5.3) ψ𝜷,𝜷¯(z)=2kImβklog1|zqk|+2kReβkarg(zqk)\psi_{\boldsymbol{\beta},\overline{\boldsymbol{\beta}}}(z)=-2\sum_{k}\mathrm{Im}\,\beta_{k}\log\frac{1}{|z-q_{k}|}+2\sum_{k}\mathrm{Re}\,\beta_{k}\arg(z-q_{k})

in the \mathbb{C}-uniformization.

For two divisors 𝝉±=τj±zj\boldsymbol{\tau}^{\pm}=\sum\tau_{j}^{\pm}\cdot z_{j} satisfying the neutrality condition (NC0),(\mathrm{NC}_{0}), we define

Ψ𝜷+,𝜷[𝝉+,𝝉]:=Ψ𝜷++[𝝉+]Ψ𝜷[𝝉],\Psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}[\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}]:=\Psi_{\boldsymbol{\beta}^{+}}^{+}[\boldsymbol{\tau}^{+}]-\Psi_{\boldsymbol{\beta}^{-}}^{-}[\boldsymbol{\tau}^{-}],

where

Ψ𝜷++[𝝉+]:=jτj+Ψ𝜷++(zj),Ψ𝜷[𝝉]:=jτjΨ𝜷(zj).\Psi_{\boldsymbol{\beta}^{+}}^{+}[\boldsymbol{\tau}^{+}]:=\sum_{j}\tau_{j}^{+}\Psi_{\boldsymbol{\beta}^{+}}^{+}(z_{j}),\qquad\Psi_{\boldsymbol{\beta}^{-}}^{-}[\boldsymbol{\tau}^{-}]:=\sum_{j}\tau_{j}^{-}\Psi_{\boldsymbol{\beta}^{-}}^{-}(z_{j}).

Then we have

Ψ𝜷+,𝜷[𝝉+,𝝉]=Ψ[𝝉+,𝝉]+ij,kτj+βk+log1zjqk+ij,kτjβklog1z¯jq¯k\Psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}[\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}]=\Psi[\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}]+i\sum_{j,k}\tau_{j}^{+}\beta_{k}^{+}\log\frac{1}{z_{j}-q_{k}}+i\sum_{j,k}\tau_{j}^{-}\beta_{k}^{-}\log\frac{1}{\bar{z}_{j}-\bar{q}_{k}}

in the \mathbb{C}-uniformization.

Sometimes it is convenient to write 𝜷=[𝜷+,𝜷],\boldsymbol{\beta}=[\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}], 𝝉=[𝝉+,𝝉],\boldsymbol{\tau}=[\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}], and

Ψ𝜷[𝝉]=Ψ𝜷+,𝜷[𝝉+,𝝉].\Psi_{\boldsymbol{\beta}}[\boldsymbol{\tau}]=\Psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}[\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}].

5.2. Stress tensors and Virasoro fields

For the reader’s convenience, we briefly review the definitions of a stress tensor and the Virasoro field. See [18, Lectures 4 and 5] for more details. Suppose A+A^{+}(A,A^{-}, respectively) is a Fock space holomorphic (anti-holomorphic, respectively) quadratic differential. Recall that a Fock space field XX is (anti-)holomorphic if the correlation z𝐄X(z)𝒳z\mapsto\mathbf{E}\,X(z)\mathcal{X} is (anti-)holomorphic in the complement of S𝒳S_{\mathcal{X}} for any tensor product 𝒳\mathcal{X} of the Gaussian free fields. Let vv be a non-random holomorphic vector field defined in some neighborhood of ζ.\zeta. We define the residue operator Av+A_{v}^{+}(AvA_{v}^{-}, respectively) as an operator on Fock space fields:

(Av+X)(z)=12πi(z)vA+X(z),(AvX)(z)=12πi(z)v¯AX(z)(A_{v}^{+}X)(z)=\frac{1}{2\pi i}\oint_{(z)}vA^{+}\,X(z),\qquad(A_{v}^{-}X)(z)=-\frac{1}{2\pi i}\oint_{(z)}\bar{v}A^{-}\,X(z)

in a given chart ϕ,ϕ(ζ)=z.\phi,\;\phi(\zeta)=z. A pair W=(A+,A)W=(A^{+},A^{-}) is called a stress tensor for a Fock space field XX if for all non-random local vector fields v,v, the so-called “residue form of Ward’s identity”

(5.4) vX=Av+X+AvX\mathcal{L}_{v}X=A^{+}_{v}X+A^{-}_{v}X

holds in the maximal open set Dhol(v)D_{\mathrm{hol}}(v) where vv is holomorphic. We recall the definition of Lie derivatives (see [18, Section 3.4]):

(vXϕ)=ddt|t=0(Xϕψt),(\mathcal{L}_{v}X\,\|\,\phi)=\frac{\mathrm{d}}{\mathrm{d}t}\Big{|}_{t=0}(X\,\|\,\phi\circ\psi_{-t}),

where ψt\psi_{t} is a local flow of v,v, and ϕ\phi is an arbitrary chart. The Lie derivative operator v\mathcal{L}_{v} depends \mathbb{R}-linearly on vector fields vv and it is convenient to consider its \mathbb{C}-linear part v+\mathcal{L}_{v}^{+} and anti-linear part v:\mathcal{L}_{v}^{-}:

2v+=viiv,2v=v+iiv.2\mathcal{L}_{v}^{+}=\mathcal{L}_{v}-i\mathcal{L}_{iv},\qquad 2\mathcal{L}_{v}^{-}=\mathcal{L}_{v}+i\mathcal{L}_{iv}.

The \mathbb{C}-linear part and the anti-linear part are related as v=v+¯\mathcal{L}_{v}^{-}=\overline{\mathcal{L}_{v}^{+}} and the conjugation means that v+¯X=v+X¯¯.\overline{\mathcal{L}_{v}^{+}}X=\overline{\mathcal{L}_{v}^{+}\bar{X}}. If XX is a (λ,λ)(\lambda,\lambda_{*})-differential, then

(5.5) v+X=(v+λv)X,vX=(v¯¯+λv¯)X;\mathcal{L}_{v}^{+}X=(v\partial+\lambda v^{\prime})X,\qquad\mathcal{L}_{v}^{-}X=(\bar{v}{\bar{\partial}}+\lambda_{*}\overline{v^{\prime}})X;

if XX is a PPS(μ+,μ)\mathrm{PPS}(\mu^{+},\mu^{-}) form, then

(5.6) v+X=vX+μ+v,vX=v¯¯X+μv¯;\mathcal{L}_{v}^{+}X=v\partial X+\mu^{+}v^{\prime},\qquad\mathcal{L}_{v}^{-}X=\bar{v}{\bar{\partial}}X+\mu^{-}\overline{v^{\prime}};

if XX is a pre-Schwarzian form of order μ,\mu, then

(5.7) vX=(v+v)X+μv′′;\mathcal{L}_{v}X=\left(v\partial+v^{\prime}\right)X+\mu v^{\prime\prime};

if XX is a Schwarzian form of order μ,\mu, then

(5.8) vX=(v+2v)X+μv′′′.\mathcal{L}_{v}X=\left(v\partial+2v^{\prime}\right)X+\mu v^{\prime\prime\prime}.

We denote by (W)(A+,A)\mathcal{F}(W)\equiv\mathcal{F}(A^{+},A^{-}) Ward’s family of W,W, the linear space of all Fock space fields XX with a stress tensor WW in common. If (A+,A)\mathcal{F}(A^{+},A^{-}) is closed under complex conjugation, we can choose A+=A,A=A¯.A^{+}=A,A^{-}=\bar{A}. In this symmetric case, the Lie derivatives operators v±\mathcal{L}^{\pm}_{v} acts on (A,A¯)\mathcal{F}(A,\bar{A}) as the residue operators Av±.A^{\pm}_{v}. In this case X(A,A¯)X\in\mathcal{F}(A,\bar{A}) if and only if the following Ward’s OPEs hold in every local chart ϕ:\phi:

(5.9) Singζz[A(ζ)X(z)]=(kζ+X)(z),Singζz[A(ζ)X¯(z)]=(kζ+X¯)(z),\operatorname{Sing}_{\zeta\to z}[A(\zeta)X(z)]=(\mathcal{L}_{k_{\zeta}}^{+}X)(z),\quad\operatorname{Sing}_{\zeta\to z}[A(\zeta)\bar{X}(z)]=(\mathcal{L}_{k_{\zeta}}^{+}\bar{X})(z),

where Singζz\operatorname{Sing}_{\zeta\to z} means the singular part of the operator product expansion in chart ϕ\phi as ζz\zeta\to z and kζk_{\zeta} is the local vector field given by

(kζϕ)(η)=1ζη.(k_{\zeta}\|\phi)(\eta)=\frac{1}{\zeta-\eta}.

See [18, Proposition 5.3]. We often use the notation \sim for the singular part of the operator product expansion.

For example, in a simply-connected domain, the Gaussian free field Φ\Phi (with Dirichlet boundary condition) has a stress tensor

(A,A¯),A=12JJ,J=Φ.(A,\bar{A}),\qquad A=-\frac{1}{2}J\odot J,\qquad J=\partial\Phi.

It follows from Ward’s OPE

A(ζ)Φ(z)Φ(z)ζz,ζz.A(\zeta)\Phi(z)\sim\frac{\partial\Phi(z)}{\zeta-z},\qquad\zeta\to z.

It is well known that Ward’s family is closed under the OPE product * and differentiations, see [18, Proposition 5.8]. This fact implies that (A,A¯)(A,\bar{A}) is also a stress tensor for JJ and T:=12JJ.T:=-\frac{1}{2}J*J.

On the Riemann sphere, the current fields J,J¯J,\bar{J} are defined as well-defined single-variable Fock space fields,

J(z)=zΨ(z,z0),J¯(z)=¯zΨ(z,z0).J(z)=\partial_{z}\Psi(z,z_{0}),\qquad\bar{J}(z)={\bar{\partial}}_{z}\Psi(z,z_{0}).

It is easy to see that this definition does not depend on the choice of z0.z_{0}. Furthermore, JJ is holomorphic and J¯\bar{J} is anti-holomorphic. As in a simply-connected domain, the Gaussian free field Ψ\Psi has a stress tensor

(A,A¯),A=12JJ.(A,\bar{A}),\qquad A=-\frac{1}{2}J\odot J.

In this case, Ward’s OPE reads as

A(ζ)Ψ(z,z0)zΨ(z,z0)ζz,ζz.A(\zeta)\Psi(z,z_{0})\sim\frac{\partial_{z}\Psi(z,z_{0})}{\zeta-z},\qquad\zeta\to z.

A pair (T+,T)(T^{+},T^{-}) of Fock space fields is called the Virasoro pair for the family (A+,A)\mathcal{F}(A^{+},A^{-}) if T±(A+,A)T^{\pm}\in\mathcal{F}(A^{+},A^{-}) and if T+A+,TA¯T^{+}-A^{+},\overline{T^{-}-A^{-}} are non-random holomorphic (or meromorphic with poles where background charges are placed, see Section 5) Schwarzian forms. In the symmetric case, TT is called the Virasoro field for Ward’s family (A,A¯).\mathcal{F}(A,\bar{A}). Both in a simply-connected domain and on the Riemann sphere, the Virasoro field TT is given by

T=12JJT=-\frac{1}{2}J*J

or by the operator product expansion

J(ζ)J(z)=1(ζz)22T(z)+o(1),ζz.J(\zeta)J(z)=-\frac{1}{(\zeta-z)^{2}}-2T(z)+o(1),\qquad\zeta\to z.

It is easy to see that TT is a Schwarzian form of order 1/12.1/12. See [18, Proposition 3.4] in a simply-connected domain or [17, Proposition 4.1] on a compact Riemann surface.

Theorem 5.2.

The bosonic field Ψ𝛃Ψ𝛃+,𝛃\Psi_{\boldsymbol{\beta}}\equiv\Psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}} has a stress tensor W𝛃(A𝛃++,A𝛃),W_{\boldsymbol{\beta}}\equiv(A_{\boldsymbol{\beta}^{+}}^{+},A_{\boldsymbol{\beta}^{-}}^{-}),

A𝜷++=A+ib+Jψ𝜷++J,A𝜷=A𝜷¯+¯(=A¯ib¯J¯ψ𝜷¯+¯J¯).A_{\boldsymbol{\beta}^{+}}^{+}=A+ib^{+}\partial J-\partial\psi_{\boldsymbol{\beta}^{+}}^{+}J,\qquad A_{\boldsymbol{\beta}^{-}}^{-}=\overline{A_{\mkern 1.5mu\overline{\mkern-1.5mu\boldsymbol{\beta}^{-}\mkern-1.5mu}\mkern 1.5mu}^{+}}(=\bar{A}-ib^{-}{\bar{\partial}}\bar{J}-\overline{\partial\psi_{\mkern 1.5mu\overline{\mkern-1.5mu\boldsymbol{\beta}^{-}\mkern-1.5mu}\mkern 1.5mu}^{+}}\bar{J}).

The Virasoro pair (T𝛃++,T𝛃)(T_{\boldsymbol{\beta}^{+}}^{+},T_{\boldsymbol{\beta}^{-}}^{-}) for the family (W𝛃)\mathcal{F}(W_{\boldsymbol{\beta}}) is given by

T𝜷++\displaystyle T_{\boldsymbol{\beta}^{+}}^{+} =12(Ψ𝜷+,𝜷Ψ𝜷+,𝜷)+ib+2Ψ𝜷+,𝜷,\displaystyle=-\frac{1}{2}(\partial\Psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}*\partial\Psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}})+ib^{+}\partial^{2}\Psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}},
T𝜷\displaystyle T_{\boldsymbol{\beta}^{-}}^{-} =12(¯Ψ𝜷+,𝜷¯Ψ𝜷+,𝜷)ib¯2Ψ𝜷+,𝜷.\displaystyle=-\frac{1}{2}({\bar{\partial}}\Psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}*{\bar{\partial}}\Psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}})-ib^{-}{\bar{\partial}}^{2}\Psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}.
Proof.

We first observe that A𝜷++A_{\boldsymbol{\beta}^{+}}^{+} is a holomorphic quadratic differential. Indeed, ib+Jib^{+}\partial J and ψ𝜷++J\partial\psi_{\boldsymbol{\beta}^{+}}^{+}J satisfy

ib+J\displaystyle ib^{+}\partial J =ib+h′′J~h+ib(h)2J~h,\displaystyle=ib^{+}h^{\prime\prime}\tilde{J}\circ h+ib(h^{\prime})^{2}\partial\tilde{J}\circ h,
ψ𝜷++J\displaystyle\partial\psi_{\boldsymbol{\beta}^{+}}^{+}J =(hψ~𝜷++h+ib+h′′h)hJ~h.\displaystyle=(h^{\prime}\partial\tilde{\psi}_{\boldsymbol{\beta}^{+}}^{+}\circ h+ib^{+}\frac{h^{\prime\prime}}{h^{\prime}})h^{\prime}\tilde{J}\circ h.

Similarly, A𝜷A_{\boldsymbol{\beta}^{-}}^{-} is an anti-holomorphic quadratic differential.

Next we check Ward’s OPE on ^\widehat{\mathbb{C}}

A𝜷++(ζ)Ψ𝜷+,𝜷(z)\displaystyle A_{\boldsymbol{\beta}^{+}}^{+}(\zeta)\Psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}(z) =(A(ζ)+ib+J(ζ)ψ𝜷++(ζ)J(ζ))(Ψ(z)+ψ𝜷++(z)+ψ𝜷+(z)¯)\displaystyle=\big{(}A(\zeta)+ib^{+}\partial J(\zeta)-\partial\psi_{\boldsymbol{\beta}^{+}}^{+}(\zeta)J(\zeta)\big{)}\big{(}\Psi(z)+\psi_{\boldsymbol{\beta}^{+}}^{+}(z)+\overline{\psi_{{\boldsymbol{\beta}^{-}}}^{+}(z)}\big{)}
J(z)ζz+ib+1(ζz)2+ψ𝜷++(z)ζz=Ψ𝜷+,𝜷(z)ζz+ib+1(ζz)2.\displaystyle\sim\frac{J(z)}{\zeta-z}+ib^{+}\frac{1}{(\zeta-z)^{2}}+\frac{\partial\psi_{\boldsymbol{\beta}^{+}}^{+}(z)}{\zeta-z}=\frac{\partial\Psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}(z)}{\zeta-z}+ib^{+}\frac{1}{(\zeta-z)^{2}}.

Similar operator product expansion holds for A𝜷(ζ)Ψ𝜷+,𝜷(z)A_{\boldsymbol{\beta}^{-}}^{-}(\zeta)\Psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}(z) as ζz.\zeta\to z.

Finally we want to show that T𝜷++T_{\boldsymbol{\beta}^{+}}^{+} is a Schwarzian form of order c+/12c^{+}/12 (c+=112(b+)2).(c^{+}=1-12(b^{+})^{2}). We find

T𝜷++=T+(A𝜷++A)12(ψ𝜷+)2+ib+2ψ𝜷+T_{\boldsymbol{\beta}^{+}}^{+}=T+(A_{\boldsymbol{\beta}^{+}}^{+}-A)-\frac{1}{2}(\partial\psi_{\boldsymbol{\beta}^{+}})^{2}+ib^{+}\partial^{2}\psi_{\boldsymbol{\beta}^{+}}

from the expressions of A𝜷++A_{\boldsymbol{\beta}^{+}}^{+} and T𝜷++.T_{\boldsymbol{\beta}^{+}}^{+}. All we need to check is that 12(ψ𝜷+)2+ib2ψ𝜷+-\frac{1}{2}(\partial\psi_{\boldsymbol{\beta}^{+}})^{2}+ib\partial^{2}\psi_{\boldsymbol{\beta}^{+}} is a Schwarzian form of order (b+)2.-(b^{+})^{2}. It follows from the transformation laws for 12(ψ𝜷+)2-\frac{1}{2}(\partial\psi_{\boldsymbol{\beta}^{+}})^{2} and ib+2ψ𝜷+:ib^{+}\partial^{2}\psi_{\boldsymbol{\beta}^{+}}:

12(ψ𝜷+)2\displaystyle-\frac{1}{2}(\partial\psi_{\boldsymbol{\beta}^{+}})^{2} =ib+h′′ψ~𝜷+h12(h)2(ψ~𝜷+h)2+12(b+)2(h′′h)2,\displaystyle=-ib^{+}h^{\prime\prime}\partial\tilde{\psi}_{\boldsymbol{\beta}^{+}}\circ h-\frac{1}{2}(h^{\prime})^{2}(\partial\tilde{\psi}_{\boldsymbol{\beta}^{+}}\circ h)^{2}+\frac{1}{2}(b^{+})^{2}\Big{(}\frac{h^{\prime\prime}}{h^{\prime}}\Big{)}^{2},
ib+2ψ𝜷+\displaystyle ib^{+}\partial^{2}\psi_{\boldsymbol{\beta}^{+}} =ib+h′′ψ~𝜷+h+ib+(h)2(2ψ~𝜷+)h(b+)2(h′′h).\displaystyle=\phantom{-}ib^{+}h^{\prime\prime}\partial\tilde{\psi}_{\boldsymbol{\beta}^{+}}\circ h+ib^{+}(h^{\prime})^{2}(\partial^{2}\tilde{\psi}_{\boldsymbol{\beta}^{+}}\,)\circ h-(b^{+})^{2}\Big{(}\frac{h^{\prime\prime}}{h^{\prime}}\Big{)}^{\prime}.

Similarly, T𝜷¯\overline{T_{\boldsymbol{\beta}^{-}}^{-}} is a Schwarzian form of order c¯/12\overline{c^{-}}/12 (c=112(b)2).(c^{-}=1-12(b^{-})^{2}).

Example.

In the ^\widehat{\mathbb{C}}-uniformization, we have

𝐄T𝜷++(z)=kλb+(βk+)(zqk)2+j<kβj+βk+(zqj)(zqk).\mathbf{E}\,T_{\boldsymbol{\beta}^{+}}^{+}(z)=\sum_{k}\frac{\lambda_{b^{+}}(\beta_{k}^{+})}{(z-q_{k})^{2}}+\sum_{j<k}\frac{\beta_{j}^{+}\beta_{k}^{+}}{(z-q_{j})(z-q_{k})}.

Denote by 𝜷(S)\mathcal{F}_{\boldsymbol{\beta}}(S^{\prime}) the OPE family of Ψ𝜷\Psi_{\boldsymbol{\beta}} on SS,S^{\prime}\subseteq S, the algebra (over \mathbb{C}) spanned by 11 and the derivatives of Ψ𝜷,\Psi_{\boldsymbol{\beta}}, 𝒪𝜷\mathcal{O}_{\boldsymbol{\beta}} with nodes in SS^{\prime} under the OPE multiplication .*. Let S:=S(supp𝜷1±supp𝜷2±).S^{*}:=S\setminus(\mathrm{supp}\,\boldsymbol{\beta}_{1}^{\pm}\cup\,\mathrm{supp}\,\boldsymbol{\beta}_{2}^{\pm}).

Theorem 5.3.

Given background charges 𝛃1±,𝛃2±\boldsymbol{\beta}^{\pm}_{1},\boldsymbol{\beta}^{\pm}_{2} with the neutrality conditions (NCb±),(\mathrm{NC}_{b^{\pm}}), the image of 𝛃1(S)\mathcal{F}_{\boldsymbol{\beta}_{1}}(S^{*}) under the insertion of eiΨ[𝛃2𝛃1]\mathrm{e}^{\odot i\Psi[\boldsymbol{\beta}_{2}-\boldsymbol{\beta}_{1}]} is 𝛃2(S)\mathcal{F}_{\boldsymbol{\beta}_{2}}(S^{*}) within correlations.

Proof.

We first show that the image of Ψ𝜷1\Psi_{\boldsymbol{\beta}_{1}} under the insertion of eiΨ[𝜷2𝜷1]\mathrm{e}^{\odot i\Psi[\boldsymbol{\beta}_{2}-\boldsymbol{\beta}_{1}]} is Ψ𝜷2\Psi_{\boldsymbol{\beta}_{2}} within correlations:

𝐄Ψ𝜷1(ζ,z)eiΨ[𝜷2𝜷1]=𝐄Ψ𝜷2(ζ,z).\mathbf{E}\,\Psi_{\boldsymbol{\beta}_{1}}(\zeta,z)\,\mathrm{e}^{\odot i\Psi[\boldsymbol{\beta}_{2}-\boldsymbol{\beta}_{1}]}=\mathbf{E}\,\Psi_{\boldsymbol{\beta}_{2}}(\zeta,z).

By Wick’s calculus, we have

(5.10) 𝐄Ψ𝜷1(ζ,z)eiΨ[𝜷2𝜷1]\displaystyle\mathbf{E}\,\Psi_{\boldsymbol{\beta}_{1}}(\zeta,z)\,\mathrm{e}^{\odot i\Psi[\boldsymbol{\beta}_{2}-\boldsymbol{\beta}_{1}]} =ψ𝜷1(ζ)ψ𝜷1(z)+𝐄Ψ(ζ,z)eiΨ[𝜷2𝜷1]\displaystyle=\psi_{\boldsymbol{\beta}_{1}}(\zeta)-\psi_{\boldsymbol{\beta}_{1}}(z)+\mathbf{E}\,\Psi(\zeta,z)\,\mathrm{e}^{\odot i\Psi[\boldsymbol{\beta}_{2}-\boldsymbol{\beta}_{1}]}
=ψ𝜷1(ζ)ψ𝜷1(z)+i𝐄Ψ(ζ,z)Ψ[𝜷2𝜷1]=ψ𝜷2(ζ)ψ𝜷2(z).\displaystyle=\psi_{\boldsymbol{\beta}_{1}}(\zeta)-\psi_{\boldsymbol{\beta}_{1}}(z)+i\,\mathbf{E}\,\Psi(\zeta,z)\Psi[\boldsymbol{\beta}_{2}-\boldsymbol{\beta}_{1}]=\psi_{\boldsymbol{\beta}_{2}}(\zeta)-\psi_{\boldsymbol{\beta}_{2}}(z).

We define the correspondence 𝒳𝒳^\mathcal{X}\mapsto\widehat{\mathcal{X}} by the formula Ψ𝜷1Ψ𝜷2\Psi_{\boldsymbol{\beta}_{1}}\mapsto\Psi_{\boldsymbol{\beta}_{2}} and the rules

𝒳𝒳^,¯𝒳¯𝒳^,α𝒳+β𝒴α𝒳^+β𝒴^,𝒳𝒴𝒳^𝒴^.\partial\mathcal{X}\mapsto\partial\widehat{\mathcal{X}},\quad{\bar{\partial}}\mathcal{X}\mapsto{\bar{\partial}}\widehat{\mathcal{X}},\quad\alpha\mathcal{X}+\beta\mathcal{Y}\mapsto\alpha\widehat{\mathcal{X}}+\beta\widehat{\mathcal{Y}},\quad\mathcal{X}\odot\mathcal{Y}\mapsto\widehat{\mathcal{X}}\odot\widehat{\mathcal{Y}}.

Denote

𝐄^𝒳=𝐄𝒳eiΨ[𝜷2𝜷1].\widehat{\mathbf{E}}\,\mathcal{X}=\mathbf{E}\,\mathcal{X}\mathrm{e}^{\odot i\Psi[\boldsymbol{\beta}_{2}-\boldsymbol{\beta}_{1}]}.

It suffices to show

𝐄^𝒳=𝐄𝒳^\widehat{\mathbf{E}}\,\mathcal{X}=\mathbf{E}\,\widehat{\mathcal{X}}

for 𝒳=X1(ζ1,z1)Xn(ζn,zn),\mathcal{X}=X_{1}(\zeta_{1},z_{1})\odot\cdots\odot X_{n}(\zeta_{n},z_{n}), Xj=ζjmj¯ζjm~jzjnj¯zjn~jΨ(ζj,zj).X_{j}=\partial_{\zeta_{j}}^{m_{j}}{\bar{\partial}}_{\zeta_{j}}^{\widetilde{m}_{j}}\partial_{z_{j}}^{n_{j}}{\bar{\partial}}_{z_{j}}^{\widetilde{n}_{j}}\Psi(\zeta_{j},z_{j}). Differentiating (5.10), it follows from Wick’s calculus that

𝐄X1(ζ1,z1)Xn(ζn,zn)Ψk[𝜷2𝜷1]={k!j=1n𝐄Xj(ζj,zj)Ψ[𝜷2𝜷1]if n=k,0otherwise.\mathbf{E}\,X_{1}(\zeta_{1},z_{1})\odot\cdots\odot X_{n}(\zeta_{n},z_{n})\Psi^{\odot k}[\boldsymbol{\beta}_{2}-\boldsymbol{\beta}_{1}]=\begin{cases}k!\prod_{j=1}^{n}\,\mathbf{E}\,X_{j}(\zeta_{j},z_{j})\Psi[\boldsymbol{\beta}_{2}-\boldsymbol{\beta}_{1}]&\textrm{if }n=k,\\ 0&\textrm{otherwise.}\end{cases}

We now compute

𝐄^𝒳\displaystyle\widehat{\mathbf{E}}\,\mathcal{X} =k=0ikk!𝐄X1(ζ1,z1)Xn(ζn,zn)Ψk[𝜷2𝜷1]\displaystyle=\sum_{k=0}^{\infty}\frac{i^{k}}{k!}\,\mathbf{E}\,X_{1}(\zeta_{1},z_{1})\odot\cdots\odot X_{n}(\zeta_{n},z_{n})\Psi^{\odot k}[\boldsymbol{\beta}_{2}-\boldsymbol{\beta}_{1}]
=j=1ni𝐄Xj(ζj,zj)Ψ[𝜷2𝜷1]=j=1n𝐄X^j(ζj,zj)=𝐄𝒳^,\displaystyle=\prod_{j=1}^{n}i\mathbf{E}\,X_{j}(\zeta_{j},z_{j})\Psi[\boldsymbol{\beta}_{2}-\boldsymbol{\beta}_{1}]=\prod_{j=1}^{n}\mathbf{E}\,\widehat{X}_{j}(\zeta_{j},z_{j})=\mathbf{E}\,\widehat{\mathcal{X}},

which completes the proof. ∎

5.3. Embeddings

In this subsection we construct the Gaussian free field in a simply-connected domain DD with Dirichlet boundary condition from the Gaussian free field on its Schottky double S=Ddouble.S=D^{\mathrm{double}}. Recall that the Gaussian free field can be defined as a Gaussian field indexed by the energy space. See Subsection 3.5 for this definition on ^.\widehat{\mathbb{C}}. The energy space (D)\mathcal{E}(D) can be embedded isometrically into (S)\mathcal{E}(S) in a natural way. For example, for μ\mu in the energy space (),\mathcal{E}(\mathbb{H}),

μ()2\displaystyle\|\mu\|_{\mathcal{E}(\mathbb{H})}^{2} =×log1|zw|2μ(z)μ(w)¯×log1|zw¯|2μ(z)μ(w)¯\displaystyle=\int_{\mathbb{H}\times\mathbb{H}}\log\frac{1}{|z-w|^{2}}\,\mu(z)\overline{\mu(w)}-\int_{\mathbb{H}\times\mathbb{H}}\log\frac{1}{|z-\bar{w}|^{2}}\,\mu(z)\overline{\mu(w)}
=12×log1|zw|2ν(z)ν(w)¯=12ν(^)2,\displaystyle=\frac{1}{2}\int_{\mathbb{C}\times\mathbb{C}}\log\frac{1}{|z-w|^{2}}\,\nu(z)\overline{\nu(w)}=\frac{1}{2}\|\nu\|_{\mathcal{E}(\widehat{\mathbb{C}})}^{2},

where

ν:={μ in ,μ in ,\nu:=\begin{cases}\phantom{-}\mu\phantom{{}^{*}}\textrm{ in }\mathbb{H},\\ -\mu^{*}\textrm{ in }\mathbb{H}^{*},\end{cases}

and E:={z¯|zE},μ(E):=μ(E).E^{*}:=\{\bar{z}\,|\,z\in E\},\mu^{*}(E):=\mu(E^{*}). As a Fock space field, the Gaussian free field ΦΦD\Phi\equiv\Phi_{D} in DD with Dirichlet boundary condition can be constructed from the Gaussian free field ΨΨS\Psi\equiv\Psi_{S} on S:S:

ΦD(z)=12ΨS(z,z),\Phi_{D}(z)=\frac{1}{\sqrt{2}}\Psi_{S}(z,z^{*}),

where ι:SS,zz\iota:S\to S,z\mapsto z^{*} is the canonical involution in the Schottky double SS of D.D. For example,

Φ(z)=12Ψ^(z,z¯),\Phi_{\mathbb{H}}(z)=\frac{1}{\sqrt{2}}\Psi_{\widehat{\mathbb{C}}}(z,\bar{z}),

and

Φ𝔻(z)=12Ψ^(z,z),z=1/z¯.\Phi_{\mathbb{D}}(z)=\frac{1}{\sqrt{2}}\Psi_{\widehat{\mathbb{C}}}(z,z^{*}),\qquad z^{*}=1/\bar{z}.

In the chordal case, the current fields and the stress energy tensors are related as

J(z)=12(J(z)J¯(z¯))J_{\mathbb{H}}(z)=\frac{1}{\sqrt{2}}\big{(}J_{\mathbb{C}}(z)-{\bar{J}_{\mathbb{C}}(\bar{z})}\big{)}

and

A(z)=12(A(z)+A¯(z¯)+J(z)J¯(z¯)).A_{\mathbb{H}}(z)=\frac{1}{2}\big{(}A_{\mathbb{C}}(z)+{\bar{A}_{\mathbb{C}}(\bar{z})}+J_{\mathbb{C}}(z)\odot{\bar{J}_{\mathbb{C}}(\bar{z})}\big{)}.

Similar statements hold in the radial case. Furthermore, the formal bosonic fields are related as

Φ±(z)=12(Ψ±(z)Ψ(z)).\Phi^{\pm}(z)=\frac{1}{\sqrt{2}}\big{(}\Psi^{\pm}(z)-\Psi^{\mp}(z^{*})\big{)}.

More generally, for a double divisor (𝝈+,𝝈)(\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}) satisfying the neutrality condition (NC0),(\mathrm{NC}_{0}), we have

(5.11) Φ[𝝈+,𝝈]=12Ψ[𝝈++𝝈,𝝈++𝝈].\Phi[\boldsymbol{\sigma}^{+},\boldsymbol{\sigma}^{-}]=\frac{1}{\sqrt{2}}\Psi[\boldsymbol{\sigma}^{+}+\boldsymbol{\sigma}_{*}^{-},\boldsymbol{\sigma}_{*}^{+}+\boldsymbol{\sigma}^{-}].

In Subsection 9.1 we construct the Gaussian free field in DD with Neumann boundary condition from ΨS.\Psi_{S}.

5.4. Modification of Gaussian free field in a simply-connected domain

In Subsection 5.1 we introduce background charge modifications Ψ𝜷+,𝜷\Psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}} of the Gaussian free field on the Riemann sphere. Here the background charges 𝜷±\boldsymbol{\beta}^{\pm} satisfy the neutrality condition (NCb±)(b±).(\mathrm{NC}_{b^{\pm}})(b^{\pm}\in\mathbb{C}). In this subsection, from Ψ𝜷+,𝜷,\Psi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}, we construct background charge modifications of the Gaussian free field with Dirichlet boundary condition in a simply-connected domain.

Let us fix a background charge parameter b.b\in\mathbb{R}. This parameter bb is related to the central charge cc as c=112b2.c=1-12b^{2}. For a double background charge (𝜷+,𝜷)(\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}) on D¯×D\mkern 1.5mu\overline{\mkern-1.5muD\mkern-1.5mu}\mkern 1.5mu\times D satisfying the neutrality condition (NCb):(\mathrm{NC}_{b}):

(𝜷++𝜷)=2b,\int(\boldsymbol{\beta}^{+}+\boldsymbol{\beta}^{-})=2b,

we set 𝜷=𝜷++𝜷\boldsymbol{\beta}=\boldsymbol{\beta}^{+}+\boldsymbol{\beta}^{-}_{*} and define

Φ𝜷(z)Φ𝜷+,𝜷(z)=12Ψ2𝜷+,2𝜷(z,z).\Phi_{\boldsymbol{\beta}}(z)\equiv\Phi_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}(z)=\frac{1}{\sqrt{2}}\Psi_{\sqrt{2}\,\boldsymbol{\beta}^{+},\sqrt{2}\,\boldsymbol{\beta}_{*}^{-}}(z,z^{*}).

Then Φ𝜷\Phi_{\boldsymbol{\beta}} is a PPS(ib,ib)\mathrm{PPS}(ib,-ib) form and

φ𝜷(z):=𝐄Φ𝜷(z)=(βk++βk)(arg(zqk)+arg(zq¯k))+i(βk+βk)log|zq¯kzqk|\varphi_{\boldsymbol{\beta}}(z):=\mathbf{E}\,\Phi_{\boldsymbol{\beta}}(z)=\sum(\beta_{k}^{+}+\beta_{k}^{-})\big{(}\arg(z-q_{k})+\arg(z-\bar{q}_{k})\big{)}+i\sum(\beta_{k}^{+}-\beta_{k}^{-})\log\Big{|}\frac{z-\bar{q}_{k}}{z-q_{k}}\Big{|}

in the \mathbb{H}-uniformization.

We also define background charge modifications of formal bosonic fields as

Φ𝜷±(z)=12(Ψ2𝜷+,2𝜷±(z)Ψ2𝜷+,2𝜷(z)).\Phi^{\pm}_{\boldsymbol{\beta}}(z)=\frac{1}{\sqrt{2}}\big{(}\Psi^{\pm}_{\sqrt{2}\boldsymbol{\beta}^{+},\sqrt{2}\boldsymbol{\beta}_{*}^{-}}(z)-\Psi^{\mp}_{\sqrt{2}\boldsymbol{\beta}^{+},\sqrt{2}\boldsymbol{\beta}_{*}^{-}}(z^{*})\big{)}.

For a double divisor (𝝉+,𝝉)(\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}) satisfying the neutrality condition (NC0),(\mathrm{NC}_{0}), we set 𝝉=𝝉++𝝉.\boldsymbol{\tau}=\boldsymbol{\tau}^{+}+\boldsymbol{\tau}_{*}^{-}. Then we have

(5.12) Φ𝜷[𝝉]=12Ψ2𝜷+,2𝜷[𝝉,𝝉].\Phi_{\boldsymbol{\beta}}[\boldsymbol{\tau}]=\frac{1}{\sqrt{2}}\Psi_{\sqrt{2}\,\boldsymbol{\beta}^{+},\sqrt{2}\,\boldsymbol{\beta}_{*}^{-}}[\boldsymbol{\tau},\boldsymbol{\tau}_{*}].

• Chordal case without spins.  For a simply-connected domain DD with marked boundary points q,qkD,q,q_{k}\in\partial D, a background charge 𝜷=(2bβk)q+βkqk,(βk),\boldsymbol{\beta}=(2b-\sum\beta_{k})\cdot q+\sum\beta_{k}\cdot q_{k},(\beta_{k}\in\mathbb{R}), and a conformal map wwD,q:(D,q)(,)w\equiv w_{D,q}:~{}(D,q)\to(\mathbb{H},\infty) we have

Φ𝜷=Φ+φ𝜷,φ𝜷=2bargw+2βkarg(ww(qk)).\Phi_{\boldsymbol{\beta}}=\Phi+\varphi_{\boldsymbol{\beta}},\qquad\varphi_{\boldsymbol{\beta}}=-2b\arg w^{\prime}+2\sum\beta_{k}\arg(w-w(q_{k})).

We also define J𝜷:=Φ𝜷.J_{\boldsymbol{\beta}}:=\partial\Phi_{\boldsymbol{\beta}}. Then it is a PS(ib)\mathrm{PS}(ib) form and

J𝜷=J+j𝜷,j𝜷=ibw′′wiβkwww(qk).J_{\boldsymbol{\beta}}=J+j_{\boldsymbol{\beta}},\qquad j_{\boldsymbol{\beta}}=ib\frac{w^{\prime\prime}}{w^{\prime}}-i\sum\beta_{k}\frac{w^{\prime}}{w-w(q_{k})}.

In particular, if 𝜷=2bq,\boldsymbol{\beta}=2b\cdot q, then we have

(5.13) Φ𝜷=Φ2bargw,J𝜷=Φ𝜷=J+ibw′′w,\Phi_{\boldsymbol{\beta}}=\Phi-2b\arg w^{\prime},\qquad J_{\boldsymbol{\beta}}=\partial\Phi_{\boldsymbol{\beta}}=J+ib\frac{w^{\prime\prime}}{w^{\prime}},

and the 1-point function φ𝜷=𝐄Φ𝜷=2bargw\varphi_{\boldsymbol{\beta}}=\mathbf{E}\,\Phi_{\boldsymbol{\beta}}=-2b\arg w^{\prime} does not depend on the choice of the conformal map. Due to this property, these modifications (5.13) are well-defined. As a PS(ib)\mathrm{PS}(ib) form, the current field J𝜷J_{\boldsymbol{\beta}} is conformally invariant with respect to Aut(D,q).\mathrm{Aut}(D,q).

• Radial case with a spin only at a marked interior point.  For a simply-connected domain DD with a marked interior point qD,q\in D, we consider a conformal map

wwD,q:(D,q)(𝔻,0),w\equiv w_{D,q}:~{}(D,q)\to(\mathbb{D},0),

from DD onto the unit disc 𝔻={z:|z|<1}.\mathbb{D}=\{z\in\mathbb{C}:|z|<1\}. Let qkDq_{k}\in\partial D be marked boundary points. Given a background charge 𝜷=βqq+βq¯q+βkqk(βk)\boldsymbol{\beta}=\beta_{q}\cdot q+\overline{\beta_{q}}\cdot q^{*}+\sum\beta_{k}\cdot q_{k}(\beta_{k}\in\mathbb{R}) with the neutrality condition (NCb):(\mathrm{NC}_{b}):

2Reβq+βk=2b,2\,\mathrm{Re}\,\beta_{q}+\sum\beta_{k}=2b,

the 1-point function φ𝜷=𝐄Φ𝜷\varphi_{\boldsymbol{\beta}}=\mathbf{E}\,\Phi_{\boldsymbol{\beta}} is given by

φ𝜷(z)=2Reβqargz+2Imβqlog|z|+2βkarg(zqk).\varphi_{\boldsymbol{\beta}}(z)=2\,\mathrm{Re}\,\beta_{q}\arg z+2\,\mathrm{Im}\,\beta_{q}\log|z|+2\sum\beta_{k}\arg(z-q_{k}).

in the 𝔻\mathbb{D}-uniformization. Also we define J𝜷:=Φ𝜷.J_{\boldsymbol{\beta}}:=\partial\Phi_{\boldsymbol{\beta}}. Then it is a PS(ib)\mathrm{PS}(ib) form and

J𝜷=J+ibw′′wiβqwwiβkwww(qk).J_{\boldsymbol{\beta}}=J+ib\frac{w^{\prime\prime}}{w^{\prime}}-i{\beta_{q}}\frac{w^{\prime}}{w}-i\sum\beta_{k}\frac{w^{\prime}}{w-w(q_{k})}.

In particular, if 𝜷=bq+bq,\boldsymbol{\beta}=b\cdot q+b\cdot q^{*}, then

(5.14) Φ𝜷=Φ+φ𝜷,φ𝜷=2bargww.\Phi_{\boldsymbol{\beta}}=\Phi+\varphi_{\boldsymbol{\beta}},\qquad\varphi_{\boldsymbol{\beta}}=-2b\arg\frac{w^{\prime}}{w}.

Since the multivalued function φ𝜷\varphi_{\boldsymbol{\beta}} does not depend on the choice of the conformal map, the “bosonic” field, Φ𝜷\Phi_{\boldsymbol{\beta}} in D:=D{q}D^{*}:=D\setminus\{q\} is invariant with respect to Aut(D,q).\mathrm{Aut}(D,q). Also we have

(5.15) J𝜷=J+j𝜷,j𝜷=φ𝜷=ib(w′′www).J_{\boldsymbol{\beta}}=J+j_{\boldsymbol{\beta}},\qquad j_{\boldsymbol{\beta}}=\partial\varphi_{\boldsymbol{\beta}}=ib\Big{(}\frac{w^{\prime\prime}}{w^{\prime}}-\frac{w^{\prime}}{w}\Big{)}.
Remarks.

(a) For b0,b\neq 0, the 1-point function φ𝜷\varphi_{\boldsymbol{\beta}} has monodromy 4πb4\pi b around q.q.

(b) For b0,b\neq 0, the 1-point function j𝜷j_{\boldsymbol{\beta}} has a simple pole at q,q, and JJ is holomorphic in D.D^{*}.

(c) As a PS(ib)\mathrm{PS}(ib) form, the current J𝜷J_{\boldsymbol{\beta}} is conformally invariant with respect to Aut(D,q).\mathrm{Aut}(D,q).

For a symmetric background charge 𝜷=βkqk+βk¯qk\boldsymbol{\beta}=\sum\beta_{k}\cdot q_{k}+\mkern 1.5mu\overline{\mkern-1.5mu\beta_{k}\mkern-1.5mu}\mkern 1.5mu\cdot q_{k}^{*} on S,S, we have

φ𝜷(z):=𝐄Φ𝜷(z)=2Reβk(arg(zqk)+arg(zq¯k))2Imβklog|zq¯kzqk|\varphi_{\boldsymbol{\beta}}(z):=\mathbf{E}\,\Phi_{\boldsymbol{\beta}}(z)=2\sum\mathrm{Re}\,\beta_{k}\big{(}\arg(z-q_{k})+\arg(z-\bar{q}_{k})\big{)}-2\sum\mathrm{Im}\,\beta_{k}\log\Big{|}\frac{z-\bar{q}_{k}}{z-q_{k}}\Big{|}

and

(5.16) j𝜷(z):=𝐄J𝜷(z)=i(βkzqk+β¯kzq¯k)j_{\boldsymbol{\beta}}(z):=\mathbf{E}\,J_{\boldsymbol{\beta}}(z)=-i\sum\Big{(}\frac{\beta_{k}}{z-q_{k}}+\frac{\bar{\beta}_{k}}{z-\bar{q}_{k}}\Big{)}

in the \mathbb{H}-uniformization if all qkq_{k}’s are in D.D. Later, we present the connection between boundary conformal field theory with symmetric background charges and the chordal/radial SLE theory with forces and spins. The following theorem is parallel with Theorem 5.2.

Theorem 5.4.

For a symmetric background charge 𝛃\boldsymbol{\beta} on S,S, the bosonic field Φ𝛃\Phi_{\boldsymbol{\beta}} in DD has a stress tensor (A𝛃,A¯𝛃):(A_{\boldsymbol{\beta}},\bar{A}_{\boldsymbol{\beta}}):

A𝜷=A+ibJj𝜷JA_{\boldsymbol{\beta}}=A+ib\partial J-j_{\boldsymbol{\beta}}J

and its Virasoro field is

T𝜷=12J𝜷J𝜷+ibJ𝜷.T_{\boldsymbol{\beta}}=-\frac{1}{2}J_{\boldsymbol{\beta}}*J_{\boldsymbol{\beta}}+ib\partial J_{\boldsymbol{\beta}}.
Example.

For a background charge 𝜷=βqq+βq¯q+βkqk(qD,qkD,βk)\boldsymbol{\beta}=\beta_{q}\cdot q+\mkern 1.5mu\overline{\mkern-1.5mu\beta_{q}\mkern-1.5mu}\mkern 1.5mu\cdot q^{*}+\sum\beta_{k}\cdot q_{k}(q\in D,q_{k}\in\partial D,\beta_{k}\in\mathbb{R}) satisfying the neutrality condition (NCb)(\mathrm{NC}_{b}), we have

𝐄T𝜷(z)=kλk(zqk)2+j<kβjβk(zqj)(zqk)+λqz2+kβqβkz(zqk)\mathbf{E}\,T_{\boldsymbol{\beta}}(z)=\sum_{k}\frac{\lambda_{k}}{(z-q_{k})^{2}}+\sum_{j<k}\frac{\beta_{j}\beta_{k}}{(z-q_{j})(z-q_{k})}+\frac{\lambda_{q}}{z^{2}}+\sum_{k}\frac{\beta_{q}\beta_{k}}{z(z-q_{k})}

in the (𝔻,0)(\mathbb{D},0)-uniformization, where λk=λb(βk)\lambda_{k}=\lambda_{b}(\beta_{k}) and λq=λb(βq).\lambda_{q}=\lambda_{b}(\beta_{q}).

5.5. OPE exponentials in a punctured domain

For simplicity, we consider the standard radial case 𝜷=bq+bq.\boldsymbol{\beta}=b\cdot q+b\cdot q^{*}. Our goal in this subsection is to explain the statement that under the neutrality condition (NC0)(\mathrm{NC}_{0}) on 𝝉=𝝉++𝝉\boldsymbol{\tau}=\boldsymbol{\tau}^{+}+\boldsymbol{\tau}^{-}_{*} the modified multi-vertex field 𝒪𝜷[𝝉]𝒪𝜷[𝝉+,𝝉]\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}]\equiv\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}] (with τq+=τq=0\tau_{q}^{+}=\tau_{q}^{-}=0) in DD^{*} can be viewed as the OPE exponential of iΦ𝜷[𝝉+,𝝉].i\Phi_{\boldsymbol{\beta}}[\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}]. Non-chiral vertex fields 𝒱𝜷(σ)\mathcal{V}_{\boldsymbol{\beta}}^{(\sigma)} are defined by

𝒱𝜷(σ):=eiσΦ𝜷=n=0(iσ)nn!Φ𝜷n.\mathcal{V}_{\boldsymbol{\beta}}^{(\sigma)}:=\mathrm{e}^{*i\sigma\Phi_{\boldsymbol{\beta}}}=\sum_{n=0}^{\infty}\frac{(i\sigma)^{n}}{n!}\,\Phi_{\boldsymbol{\beta}}^{*n}.

We have

𝒱𝜷(σ)=eiσφ𝜷𝒱(σ)=eiσφ𝜷Cσ2eiσΦ,(φ𝜷:=𝐄Φ𝜷).\mathcal{V}_{\boldsymbol{\beta}}^{(\sigma)}=\mathrm{e}^{i\sigma\varphi_{\boldsymbol{\beta}}}\mathcal{V}^{(\sigma)}=\mathrm{e}^{i\sigma\varphi_{\boldsymbol{\beta}}}C^{-\sigma^{2}}\mathrm{e}^{\odot i\sigma\Phi},\qquad(\varphi_{\boldsymbol{\beta}}:=\mathbf{E}\,\Phi_{\boldsymbol{\beta}}).

The non-random field φ𝜷\varphi_{\boldsymbol{\beta}} is a PPS(ib,ib)\mathrm{PPS}(ib,-ib) form and CC is the conformal radius. In terms of a conformal map w:(D,q)(𝔻,0),w:(D,q)\to(\mathbb{D},0), we have

(CidD)(z)=1|w(z)|2|w(z)|,(C\,\|\,\mathrm{id}_{D})(z)=\frac{1-|w(z)|^{2}}{|w^{\prime}(z)|},

where idD\mathrm{id}_{D} is the identity chart of D.D. Thus the conformal radius is a [1/2,1/2][-1/2,-1/2]-differential and 𝒱(σ)\mathcal{V}^{(\sigma)} is a [λ,λ][\lambda,\lambda_{*}]-differential with λ=λb(σ),λ=λb(σ).\lambda=\lambda_{b}(\sigma),\lambda_{*}=\lambda_{b}(-\sigma). Its 1-point function in the (𝔻,0)(\mathbb{D},0)-uniformization is

𝐄𝒱𝜷(σ)(z)=(1|z|2)σ2zσb(z¯)σb.\mathbf{E}\,\mathcal{V}_{\boldsymbol{\beta}}^{(\sigma)}(z)=(1-|z|^{2})^{-\sigma^{2}}z^{\sigma b}(\bar{z})^{-\sigma b}.

Since 𝒱𝜷(σ)(z)\mathcal{V}_{\boldsymbol{\beta}}^{(\sigma)}(z) and 𝒪𝜷[σz,σz]\mathcal{O}_{\boldsymbol{\beta}}[\sigma\cdot z,-\sigma\cdot z] contain the same Wick’s exponential and they are differentials with the same conformal dimensions and the same correlation functions in the (𝔻,0)(\mathbb{D},0)-uniformization, we conclude that

𝒪𝜷[σz,σz]=eiσΦ𝜷(z).\mathcal{O}_{\boldsymbol{\beta}}[\sigma\cdot z,-\sigma\cdot z]=\mathrm{e}^{*i\sigma\Phi_{\boldsymbol{\beta}}(z)}.

Next, we explain the OPE product of OPE exponentials and why it corresponds to the addition of divisors. Here is the typical example:

(5.17) 𝒱𝜷(σ1)𝒱𝜷(σ2)=𝒱𝜷(σ1+σ2).\mathcal{V}_{\boldsymbol{\beta}}^{(\sigma_{1})}*\mathcal{V}_{\boldsymbol{\beta}}^{(\sigma_{2})}=\mathcal{V}_{\boldsymbol{\beta}}^{(\sigma_{1}+\sigma_{2})}.

Here * means the coefficients of the leading term in OPE expansion. An alternative notation for such OPE multiplication is 𝒱𝜷(σ1)𝒱𝜷(σ2),\mathcal{V}_{\boldsymbol{\beta}}^{(\sigma_{1})}\star\mathcal{V}_{\boldsymbol{\beta}}^{(\sigma_{2})}, see [18, Section 15.2]. To see (5.17), we need to compute asymptotic behavior of exp(σ1σ2𝐄[Φ(ζ)Φ(z)]):\exp(-\sigma_{1}\sigma_{2}\mathbf{E}[\Phi(\zeta)\Phi(z)]):

|w(ζ)w(z)1w(ζ)w(z)¯|2σ1σ2|w(z)(ζz)1|w(z)|2|2σ1σ2,ζz.\bigg{|}\frac{w(\zeta)-w(z)}{1-w(\zeta)\overline{w(z)}}\bigg{|}^{2\sigma_{1}\sigma_{2}}\sim\bigg{|}\frac{w^{\prime}(z)(\zeta-z)}{1-|w(z)|^{2}}\bigg{|}^{2\sigma_{1}\sigma_{2}},\qquad\zeta\to z.

This implies eiσ1Φeiσ2Φ=C2σ1σ2ei(σ1+σ2)Φ.\mathrm{e}^{\odot i\sigma_{1}\Phi}*\mathrm{e}^{\odot i\sigma_{2}\Phi}=C^{-2\sigma_{1}\sigma_{2}}\mathrm{e}^{\odot i(\sigma_{1}+\sigma_{2})\Phi}. The same argument shows that if f1f_{1} and f2f_{2} are non-random fields, then

(f1eiσ1Φ)(f2eiσ2Φ)=f1f2C2σ1σ2ei(σ1+σ2)Φ.(f_{1}\mathrm{e}^{\odot i\sigma_{1}\Phi})*(f_{2}\mathrm{e}^{\odot i\sigma_{2}\Phi})=f_{1}f_{2}C^{-2\sigma_{1}\sigma_{2}}\mathrm{e}^{\odot i(\sigma_{1}+\sigma_{2})\Phi}.

Taking fj=𝐄𝒱𝜷(σj)=eiσjφ𝜷Cσj2,f_{j}=\mathbf{E}\,\mathcal{V}_{\boldsymbol{\beta}}^{(\sigma_{j})}=\mathrm{e}^{i\sigma_{j}\varphi_{\boldsymbol{\beta}}}C^{-\sigma_{j}^{2}}, the above identity shows (5.17).

The computation is more transparent if we use formal OPE exponentials. Let us formally define

𝒪(σ)(``"eiσΦ𝜷+)=(w)λwσbeiσΦ+\mathcal{O}^{(\sigma)}(``\equiv"\mathrm{e}^{*i\sigma\Phi_{\boldsymbol{\beta}}^{+}})=(w^{\prime})^{\lambda}w^{\sigma b}\mathrm{e}^{\odot i\sigma\Phi^{+}}

(which depends on the choice of conformal map) so that

𝒪[σzσz0]=𝒪(σ)(z)𝒪(σ)(z0)\mathcal{O}[\sigma\cdot z-\sigma\cdot z_{0}]=\mathcal{O}^{(\sigma)}(z)\mathcal{O}^{(-\sigma)}(z_{0})

holds (formally) within formal correlations. In addition to the product of 1-point functions (and Wick’s exponential), we have the interaction term

eσ2𝐄[Φ+(z)Φ+(z0)]=(w(z)w(z0))σ2\mathrm{e}^{\sigma^{2}\mathbf{E}\,[\Phi^{+}(z)\Phi^{+}(z_{0})]}=(w(z)-w(z_{0}))^{-\sigma^{2}}

in 𝒪(σ)(z)𝒪(σ)(z0)\mathcal{O}^{(\sigma)}(z)\mathcal{O}^{(-\sigma)}(z_{0}). As in the OPE product of non-chiral vertex fields, we have (again formally)

(5.18) 𝒪(σ1)𝒪(σ2)=𝒪(σ1+σ2),\mathcal{O}^{(\sigma_{1})}*\mathcal{O}^{(\sigma_{2})}=\mathcal{O}^{(\sigma_{1}+\sigma_{2})},

where the OPE product * could be further specified as the \star product (leading coefficient) or as σ1σ2*_{\sigma_{1}\sigma_{2}} (coefficient of (ζz)σ1σ2(\zeta-z)^{\sigma_{1}\sigma_{2}} in the operator product expansion of 𝒪(σ1)(ζ)𝒪(σ2)(z)\mathcal{O}^{(\sigma_{1})}(\zeta)\mathcal{O}^{(\sigma_{2})}(z)). Let us compute (5.18):

𝐄𝒪(σ1)(ζ)𝒪(σ2)(z)\displaystyle\mathbf{E}\,\mathcal{O}^{(\sigma_{1})}(\zeta)\mathcal{O}^{(\sigma_{2})}(z) =(w(ζ)w(z))σ1σ2w(ζ)λb(σ1)w(ζ)σ1bw(z)λb(σ2)w(z)σ2b\displaystyle=(w(\zeta)-w(z))^{\sigma_{1}\sigma_{2}}w^{\prime}(\zeta)^{\lambda_{b}(\sigma_{1})}w(\zeta)^{\sigma_{1}b}w^{\prime}(z)^{\lambda_{b}(\sigma_{2})}w(z)^{\sigma_{2}b}
(ζz)σ1σ2w(z)σ1σ2+λb(σ1)+λb(σ2)w(z)(σ1+σ2)b.\displaystyle\sim(\zeta-z)^{\sigma_{1}\sigma_{2}}w^{\prime}(z)^{\sigma_{1}\sigma_{2}+\lambda_{b}(\sigma_{1})+\lambda_{b}(\sigma_{2})}w(z)^{(\sigma_{1}+\sigma_{2})b}.

It is trivial to compute Wick’s exponential part in the operator product expansion. The identity λb(σ1+σ2)=λb(σ1)+λb(σ2)+σ1σ2\lambda_{b}(\sigma_{1}+\sigma_{2})=\lambda_{b}(\sigma_{1})+\lambda_{b}(\sigma_{2})+\sigma_{1}\sigma_{2} now shows (5.18).

Note that the last OPE has the form

(5.19) X(ζ)Y(z)=(ζz)αCn(z)(ζz)n,Cn:=Xα+nY.X(\zeta)Y(z)=(\zeta-z)^{\alpha}\sum{C_{n}(z)}{(\zeta-z)^{n}},\qquad C_{n}:=X*_{\alpha+n}Y.

5.6. Insertions and changes of background charges

Denote by 𝜷(D)\mathcal{F}_{\boldsymbol{\beta}}(D^{\prime}) the OPE family of Φ𝜷\Phi_{\boldsymbol{\beta}} on DD,D^{\prime}\subseteq D, the algebra (over \mathbb{C}) spanned by 11 and the derivatives of Φ𝜷,\Phi_{\boldsymbol{\beta}}, 𝒪𝜷\mathcal{O}_{\boldsymbol{\beta}} with nodes in DD^{\prime} under the OPE multiplication .*. Let D:=D(supp𝜷1±supp𝜷2±).D^{*}:=D\setminus(\mathrm{supp}\,\boldsymbol{\beta}_{1}^{\pm}\cup\,\mathrm{supp}\,\boldsymbol{\beta}_{2}^{\pm}).

Theorem 5.5.

Given background charges 𝛃1,𝛃2\boldsymbol{\beta}_{1},\boldsymbol{\beta}_{2} on SS with the neutrality conditions (NCb),(\mathrm{NC}_{b}), the image of 𝛃1(D)\mathcal{F}_{\boldsymbol{\beta}_{1}}(D^{*}) under the insertion of eiΦ[𝛃2𝛃1]e^{\odot i\Phi[\boldsymbol{\beta}_{2}-\boldsymbol{\beta}_{1}]} is 𝛃2(D)\mathcal{F}_{\boldsymbol{\beta}_{2}}(D^{*}) within correlations.

Theorem 5.5 follows immediately from Theorem 5.3 and the Schottky double construction (5.12) of Φ𝜷\Phi_{\boldsymbol{\beta}} from Ψ2𝜷+,2𝜷.\Psi_{\sqrt{2}\,\boldsymbol{\beta}^{+},\sqrt{2}\,\boldsymbol{\beta}_{*}^{-}}.

We now present a typical example of the insertion operators which create the chordal/radial SLE curves. In the radial case with 𝜷ˇ=bq+bq,(qD),\check{\boldsymbol{\beta}\,}=b\cdot q+b\cdot q^{*},(q\in D), the insertion of one-leg operator

ΛpΛ(p):=𝒪𝜷ˇ[𝝉],𝝉=apa2qa2q,(pD)\Lambda_{p}\equiv\Lambda(p):=\mathcal{O}_{\check{\boldsymbol{\beta}\,}}[\boldsymbol{\tau}],\quad\boldsymbol{\tau}=a\cdot p-\frac{a}{2}\cdot q-\frac{a}{2}\cdot q^{*},\quad(p\in\partial D)

produces an operator

𝒳𝒳^\mathcal{X}\mapsto\widehat{\mathcal{X}}

acting on Fock space functionals/fields by the rules

(5.20) 𝒳𝒳^,¯𝒳¯𝒳^,α𝒳+β𝒴α𝒳^+β𝒴^,𝒳𝒴𝒳^𝒴^\partial\mathcal{X}\mapsto\partial\widehat{\mathcal{X}},\quad{\bar{\partial}}\mathcal{X}\mapsto{\bar{\partial}}\widehat{\mathcal{X}},\quad\alpha\mathcal{X}+\beta\mathcal{Y}\mapsto\alpha\widehat{\mathcal{X}}+\beta\widehat{\mathcal{Y}},\quad\mathcal{X}\odot\mathcal{Y}\mapsto\widehat{\mathcal{X}}\odot\widehat{\mathcal{Y}}

for Fock space functionals 𝒳\mathcal{X} and 𝒴\mathcal{Y} in DD^{*} and the formula

(5.21) Φ^𝜷ˇ(z)=Φ𝜷(z)=Φ𝜷ˇ(z)+aarg(1w(z))2w(z),\widehat{\Phi}_{\check{\boldsymbol{\beta}\,}}(z)=\Phi_{\boldsymbol{\beta}}(z)=\Phi_{\check{\boldsymbol{\beta}\,}}(z)+a\arg\frac{(1-w(z))^{2}}{w(z)},

where w:(D,p,q)(𝔻,1,0)w:(D,p,q)\to(\mathbb{D},1,0) is a conformal map and

(5.22) 𝜷=𝜷ˇ+𝝉=ap+(ba2)q+(ba2)q.\boldsymbol{\beta}=\check{\boldsymbol{\beta}\,}+\boldsymbol{\tau}=a\cdot p+\Big{(}b-\frac{a}{2}\Big{)}\cdot q+\Big{(}b-\frac{a}{2}\Big{)}\cdot q^{*}.

In the chordal case with 𝜷ˇ=2bq,(qD),\check{\boldsymbol{\beta}\,}=2b\cdot q,(q\in\partial D), recall that the insertion of one-leg operator

ΛpΛ(p):=𝒪𝜷ˇ[𝝉],𝝉=apaq,(pD)\Lambda_{p}\equiv\Lambda(p):=\mathcal{O}_{\check{\boldsymbol{\beta}\,}}[\boldsymbol{\tau}],\quad\boldsymbol{\tau}=a\cdot p-a\cdot q,\quad(p\in\partial D)

produces an operator 𝒳𝒳^\mathcal{X}\mapsto\widehat{\mathcal{X}} by the rules (5.20) and the formula

Φ^𝜷ˇ(z)=Φ𝜷(z)=Φ𝜷ˇ(z)+2aargw(z),\widehat{\Phi}_{\check{\boldsymbol{\beta}\,}}(z)=\Phi_{\boldsymbol{\beta}}(z)=\Phi_{\check{\boldsymbol{\beta}\,}}(z)+2a\arg w(z),

where w:(D,p,q)(,0,)w:(D,p,q)\to(\mathbb{H},0,\infty) is a conformal map and

𝜷=𝜷ˇ+𝝉=ap+(2ba)q.\boldsymbol{\beta}=\check{\boldsymbol{\beta}\,}+\boldsymbol{\tau}=a\cdot p+(2b-a)\cdot q.

See [18, Section 14.2].

Let

𝐄^[𝒳]:=𝐄[Λp𝒳]𝐄[Λp]=𝐄[eiΦ[𝝉]𝒳],𝝉=apa2qa2q.\widehat{\mathbf{E}}[\mathcal{X}]:=\frac{\mathbf{E}[\Lambda_{p}\mathcal{X}]}{\mathbf{E}[\Lambda_{p}]}=\mathbf{E}[\mathrm{e}^{\odot i\Phi[\boldsymbol{\tau}]}\mathcal{X}],\quad\boldsymbol{\tau}=a\cdot p-\frac{a}{2}\cdot q-\frac{a}{2}\cdot q^{*}.

Let 𝒳^𝜷(D)\widehat{\mathcal{X}}\in\mathcal{F}_{\boldsymbol{\beta}}(D^{*}) correspond to the string 𝒳𝜷ˇ(D)\mathcal{X}\in\mathcal{F}_{\check{\boldsymbol{\beta}}\,}(D^{*}) under the map given by (5.20) – (5.21). Then by Theorem 5.5 we have

(5.23) 𝐄^[𝒳]=𝐄[𝒳^].\widehat{\mathbf{E}}[\mathcal{X}]=\mathbf{E}[\widehat{\mathcal{X}}].
Examples.

Let φ𝜷:=𝐄Φ𝜷\varphi_{\boldsymbol{\beta}}:=\mathbf{E}\,\Phi_{\boldsymbol{\beta}} in the radial case with 𝜷\boldsymbol{\beta} in (5.22).

(a) The current J𝜷J_{\boldsymbol{\beta}} is a pre-Schwarzian form of order ib,ib,

(5.24) J𝜷=J+iaw1w+(ia2ib)ww+ibw′′w.J_{\boldsymbol{\beta}}=J+ia\frac{w^{\prime}}{1-w}+\Big{(}\frac{ia}{2}-ib\Big{)}\frac{w^{\prime}}{w}+ib\frac{w^{\prime\prime}}{w^{\prime}}.

In the (𝔻,1,0)(\mathbb{D},1,0)-uniformization,

j𝜷(z)=𝐄J𝜷(z)=ia11z+(ia2ib)1z;j_{\boldsymbol{\beta}}(z)=\mathbf{E}\,J_{\boldsymbol{\beta}}(z)=ia\dfrac{1}{1-z}+\Big{(}\dfrac{ia}{2}-ib\Big{)}\dfrac{1}{z};

(b) The Virasoro field T𝜷T_{\boldsymbol{\beta}} is a Schwarzian form of order 112c,\frac{1}{12}c,

(5.25) T𝜷\displaystyle T_{\boldsymbol{\beta}} =12J𝜷J𝜷+ibJ𝜷\displaystyle=-\dfrac{1}{2}J_{\boldsymbol{\beta}}*J_{\boldsymbol{\beta}}+ib\partial J_{\boldsymbol{\beta}}
=Aj𝜷J+ibJ+c12Sw+h1,2w2w(1w)2+h0,1/2w2w2(A=12JJ),\displaystyle=A-j_{\boldsymbol{\beta}}J+ib\partial J+\frac{c}{12}S_{w}+h_{1,2}\frac{w^{\prime 2}}{w(1-w)^{2}}+h_{0,1/2}\frac{w^{\prime 2}}{w^{2}}\qquad(A=-\frac{1}{2}J\odot J),

where h1,2=12a2ab,h_{1,2}=\frac{1}{2}a^{2}-ab, h0,1/2=18a212b2,h_{0,1/2}=\frac{1}{8}{a^{2}}-\frac{1}{2}b^{2}, (see (4.1) and (4.2)) and Sw=Nw12Nw2,(Nw=(logw))S_{w}=N_{w}^{\prime}-\frac{1}{2}{N_{w}^{2}},\,(N_{w}=(\log w^{\prime})^{\prime}) is the Schwarzian derivative of w.w. In the (𝔻,1,0)(\mathbb{D},1,0)-uniformization,

𝐄T𝜷(z)=h1,2z(1z)2+h0,1/2z2;\mathbf{E}\,T_{\boldsymbol{\beta}}(z)=\dfrac{h_{1,2}}{z(1-z)^{2}}+\dfrac{h_{0,1/2}}{z^{2}};

(c) The operator produced by the insertion of one-leg operator can be extended to the formal fields. For example, we have

Φ𝜷+=Φ𝜷ˇ++ia2logwialog(1w);\Phi_{\boldsymbol{\beta}}^{+}=\Phi_{\check{\boldsymbol{\beta}\,}}^{+}+\dfrac{ia}{2}\log w-ia\log(1-w);

and

𝒪𝜷(σ)=(1ww)aσ𝒪𝜷ˇ(σ).\mathcal{O}_{\boldsymbol{\beta}}^{(\sigma)}=\Big{(}\frac{1-w}{\sqrt{w}}\Big{)}^{a\sigma}\mathcal{O}_{\check{\boldsymbol{\beta}\,}}^{(\sigma)}.

6. Extended OPE family

In this section we extend the OPE family 𝜷(D)\mathcal{F}_{\boldsymbol{\beta}}(D^{*}) (D:=D(supp𝜷+supp𝜷)D^{*}:=D\setminus(\mathrm{supp}\,\boldsymbol{\beta}^{+}\cup\,\mathrm{supp}\,\boldsymbol{\beta}^{-}), 𝜷=𝜷++𝜷\boldsymbol{\beta}=\boldsymbol{\beta}^{+}+\boldsymbol{\beta}^{-}_{*}) of Φ𝜷\Phi_{\boldsymbol{\beta}} by adding the generators obtained from the operator product expansion of fields in 𝜷(D)\mathcal{F}_{\boldsymbol{\beta}}(D^{*}) at the punctures qksupp𝜷+supp𝜷q_{k}\in\mathrm{supp}\,\boldsymbol{\beta}^{+}\cup\,\mathrm{supp}\,\boldsymbol{\beta}^{-} or the rooting procedure. Examples include the OPE exponentials of Φ𝜷\Phi_{\boldsymbol{\beta}} with nodes at qksupp𝜷+supp𝜷.q_{k}\in\mathrm{supp}\,\boldsymbol{\beta}^{+}\cup\,\mathrm{supp}\,\boldsymbol{\beta}^{-}.

6.1. OPE at the puncture

Let us consider the radial case first. For simplicity we only consider the “formal” holomorphic puncture differential in the case that 𝜷ˇ=bq+bq:\check{\boldsymbol{\beta}\,}=b\cdot q+b\cdot q^{*}:

𝒪q(τ)=(wq)τ2/2eiτΦ+(q).\mathcal{O}_{q}^{(\tau)}=(w_{q}^{\prime})^{\tau^{2}/2}\mathrm{e}^{\odot i\tau\Phi^{+}(q)}.

This is a (formal) Fock space correlation functional with conformal dimension hq=τq2/2h_{q}=\tau_{q}^{2}/2 at q.q.

Proposition 6.1.

We have

𝒪(σ)(z)𝒪q(τ)(zq)σ(τ+b)𝒪q(σ+τ).\mathcal{O}^{(\sigma)}(z)\mathcal{O}_{q}^{(\tau)}\sim(z-q)^{\sigma(\tau+b)}\mathcal{O}_{q}^{(\sigma+\tau)}.
Proof.

As zq,z\to q, we have

𝒪(σ)(z)𝒪q(τ)\displaystyle\mathcal{O}^{(\sigma)}(z)\mathcal{O}_{q}^{(\tau)} =wσb(w)λb(σ)eiσΦ+(z)(wq)τ2/2eiτΦ+(q)\displaystyle=w^{\sigma b}(w^{\prime})^{\lambda_{b}(\sigma)}\mathrm{e}^{\odot i\sigma\Phi^{+}(z)}(w_{q}^{\prime})^{\tau^{2}/2}\mathrm{e}^{\odot i\tau\Phi^{+}(q)}
=wσ(τ+b)(w)λb(σ)(wq)τ2/2eiσΦ+(z)+iτΦ+(q)\displaystyle=w^{\sigma(\tau+b)}(w^{\prime})^{\lambda_{b}(\sigma)}(w_{q}^{\prime})^{\tau^{2}/2}\mathrm{e}^{\odot i\sigma\Phi^{+}(z)+i\tau\Phi^{+}(q)}
(zq)σ(τ+b)(wq)σ(τ+b)+λb(σ)+τ2/2ei(σ+τ)Φ+(q)\displaystyle\sim(z-q)^{\sigma(\tau+b)}(w_{q}^{\prime})^{\sigma(\tau+b)+\lambda_{b}(\sigma)+\tau^{2}/2}\mathrm{e}^{\odot i(\sigma+\tau)\Phi^{+}(q)}
=(zq)σ(τ+b)𝒪q(σ+τ).\displaystyle=(z-q)^{\sigma(\tau+b)}\mathcal{O}_{q}^{(\sigma+\tau)}.

Here we use eiσΦ+(z)eiτΦ+(q)=wστei(σΦ+(z)+τΦ+(q)),\mathrm{e}^{\odot i\sigma\Phi^{+}(z)}\mathrm{e}^{\odot i\tau\Phi^{+}(q)}=w^{\sigma\tau}\mathrm{e}^{\odot i(\sigma\Phi^{+}(z)+\tau\Phi^{+}(q))}, w(z)(zq)wqw(z)\sim(z-q)w_{q}^{\prime} as zq,z\to q, and the identity λb(σ)+τ2/2+σ(τ+b)=(σ+τ)2/2.\lambda_{b}(\sigma)+\tau^{2}/2+\sigma(\tau+b)=(\sigma+\tau)^{2}/2. The notation \sim means the first leading term of the operator product expansion. ∎

We can express the above proposition in any of the following formulas:

𝒪(σ)𝒪q(τ)𝒪(σ)σ(τ+b)𝒪q(τ)=𝒪q(σ+τ),𝒪(σ)𝒪q(τ)=𝒪q(σ+τ).\mathcal{O}^{(\sigma)}*\mathcal{O}_{q}^{(\tau)}\equiv\mathcal{O}^{(\sigma)}*_{\sigma(\tau+b)}\mathcal{O}_{q}^{(\tau)}=\mathcal{O}_{q}^{(\sigma+\tau)},\qquad\mathcal{O}^{(\sigma)}\star\mathcal{O}_{q}^{(\tau)}=\mathcal{O}_{q}^{(\sigma+\tau)}.

The point is that the arithmetic of divisors has the OPE nature, both in D:=D{q}D^{*}:=D\setminus\{q\} and at the puncture q.q.

Let us mention a special case (τ=0\tau=0) of the proposition.

Corollary 6.2.

We have

𝒪q(σ)=𝒪(σ)1q.\mathcal{O}_{q}^{(\sigma)}=\mathcal{O}^{(\sigma)}*1_{q}.

This OPE 𝒪(σ)(z)(zq)σb𝒪q(σ)(zq)\mathcal{O}^{(\sigma)}(z)\sim(z-q)^{\sigma b}\mathcal{O}_{q}^{(\sigma)}\,(z\to q) is related to the “rooting” procedure (see [18, Section 12.3] for this procedure in the chordal case):

𝒪q(σ)=limε0zεσb𝒪(σ)(zε),\mathcal{O}_{q}^{(\sigma)}=\lim_{\varepsilon\to 0}z_{\varepsilon}^{-\sigma b}\,\mathcal{O}^{(\sigma)}(z_{\varepsilon}),

where the point zεz_{\varepsilon} is at distance ε\varepsilon from q=0q=0 in the chart. Equivalently, this definition can be reached by applying the following rooting rule to the formal field 𝒪(σ)(z)=w(z)σbw(z)λeiσΦ(z):\mathcal{O}^{(\sigma)}(z)=w(z)^{\sigma b}w^{\prime}(z)^{\lambda}\,\mathrm{e}^{\odot i\sigma\Phi(z)}:

  1. (a)

    the point zz in the terms w(z)λw^{\prime}(z)^{\lambda} and eiσΦ(z)\mathrm{e}^{\odot i\sigma\Phi(z)} is replaced by the puncture q;q;

  2. (b)

    the term w(z)σbw(z)^{\sigma b} is replaced by (wq)σb.(w_{q}^{\prime})^{\sigma b}.

In the chordal case with 𝜷ˇ=2bq,\check{\boldsymbol{\beta}\,}=2b\cdot q, we consider the “formal” boundary puncture differential

𝒪q(τ)=(wq)λb(τ)eiτΦ+(q).\mathcal{O}_{q}^{(\tau)}=(w_{q}^{\prime})^{\lambda_{b}(-\tau)}\mathrm{e}^{\odot i\tau\Phi^{+}(q).}

In terms of a uniformizing map w:(D,q)(,0)w:(D,q)\to(\mathbb{H},0) we have 𝒪(σ)=(w)λb(σ)w2σb\mathcal{O}^{(\sigma)}=(w^{\prime})^{\lambda_{b}(\sigma)}w^{2\sigma b} and

𝒪(σ)(z)𝒪q(τ)(zq)σ(τ+2b)𝒪q(σ+τ)\mathcal{O}^{(\sigma)}(z)\mathcal{O}_{q}^{(\tau)}\sim(z-q)^{\sigma(\tau+2b)}\mathcal{O}_{q}^{(\sigma+\tau)}

as zq.z\to q. We use the identity λb(σ)+λb(τ)+σ(τ+2b)=λb(στ).\lambda_{b}(\sigma)+\lambda_{b}(-\tau)+\sigma(\tau+2b)=\lambda_{b}(-\sigma-\tau). This OPE can be expressed as

𝒪(σ)𝒪q(τ)𝒪(σ)σ(τ+2b)𝒪q(τ)=𝒪q(σ+τ),𝒪(σ)𝒪q(τ)=𝒪q(σ+τ).\mathcal{O}^{(\sigma)}*\mathcal{O}_{q}^{(\tau)}\equiv\mathcal{O}^{(\sigma)}*_{\sigma(\tau+2b)}\mathcal{O}_{q}^{(\tau)}=\mathcal{O}_{q}^{(\sigma+\tau)},\qquad\mathcal{O}^{(\sigma)}\star\mathcal{O}_{q}^{(\tau)}=\mathcal{O}_{q}^{(\sigma+\tau)}.

We refer to (5.19) to remind the readers what this type of OPE is. In the special case τ=0,\tau=0, we have

𝒪q(σ)=𝒪(σ)2σb1q.\mathcal{O}_{q}^{(\sigma)}=\mathcal{O}^{(\sigma)}*_{2\sigma b}1_{q}.

This is related to the “rooting” procedure:

𝒪q(σ)=limε0zε2σb𝒪(σ)(zε),\mathcal{O}_{q}^{(\sigma)}=\lim_{\varepsilon\to 0}z_{\varepsilon}^{-2\sigma b}\mathcal{O}^{(\sigma)}(z_{\varepsilon}),

where the point zεz_{\varepsilon} is at distance ε\varepsilon from q=0q=0 in the chart. In terms of a uniformizing map w:(D,q)(,)w:(D,q)\to(\mathbb{H},\infty) satisfying w(ζ)1/(ζq)w(\zeta)\sim-1/(\zeta-q) as ζq\zeta\to q in a fixed boundary chart at q,q, then the rooting procedure becomes

𝒪q(σ)=limε0w(zε)σb𝒪(σ)(zε),\mathcal{O}_{q}^{(\sigma)}=\lim_{\varepsilon\to 0}w^{\prime}(z_{\varepsilon})^{\sigma b}\mathcal{O}^{(\sigma)}(z_{\varepsilon}),

where the point zεz_{\varepsilon} is at (spherical) distance ε\varepsilon from qq in the chart ϕ,\phi, see [18, Section 12.3].

6.2. Definition of the extended OPE family

For simplicity, we consider the standard radial case with 𝜷=bq+bq.\boldsymbol{\beta}=b\cdot q+b\cdot q^{*}. If X𝜷(D),X\in\mathcal{F}_{\boldsymbol{\beta}}(D^{*}), (D:=D{q}),(D^{*}:=D\setminus\{q\}), then for all points zDz\in D and all holomorphic (local) vector fields vv we have

(6.1) 12πi(z)vT𝜷X(z)=v+X(z),12πi(z)vT𝜷X¯(z)=v+X¯(z).\frac{1}{2\pi i}\oint_{(z)}vT_{\boldsymbol{\beta}}\,X(z)=\mathcal{L}_{v}^{+}X(z),\qquad\frac{1}{2\pi i}\oint_{(z)}vT_{\boldsymbol{\beta}}\,\bar{X}(z)=\mathcal{L}_{v}^{+}\bar{X}(z).

This is not a characterization of fields in 𝜷(D),\mathcal{F}_{\boldsymbol{\beta}}(D^{*}), e.g., XΦ(q)X\Phi(q) would satisfy (6.1). Note that the fields in 𝜷(D)\mathcal{F}_{\boldsymbol{\beta}}(D^{*}) are typically not defined at q,q, e.g., Φ𝜷,J𝜷,\Phi_{\boldsymbol{\beta}},J_{\boldsymbol{\beta}}, and T𝜷.T_{\boldsymbol{\beta}}.

A Fock space functional at qq is the value of a Fock space field at q,q,

Xq=X(q).X_{q}=X(q).

Here Fock space means the Fock space of the Gaussian free field Φ\Phi and the functionals that depend on the chart at the puncture q.q. A Fock space functional at qq correlates with Fock space strings 𝒳\mathcal{X} in D.D^{*}. In fact we can consider the (tensor) products 𝒳Xq.\mathcal{X}X_{q}. Such products are well-defined Fock space functionals. Examples of Fock space functional at qq include 1q,Φ(q),1_{q},\Phi(q), and even formal functionals 𝒪(σ)(q)\mathcal{O}^{(\sigma)}(q) (with our usual proviso of neutrality).

Definition.

We say Xq𝜷(q)X_{q}\in\mathcal{F}_{\boldsymbol{\beta}}(q) if Xq=X1((Xn1q))X_{q}=X_{1}*(\cdots*(X_{n}*1_{q})\cdots) for some fields X1,Xn𝜷(D).X_{1},\cdots X_{n}\in\mathcal{F}_{\boldsymbol{\beta}}(D^{*}). The (extended) OPE family 𝜷=𝜷(D,q)\mathcal{F}_{\boldsymbol{\beta}}=\mathcal{F}_{\boldsymbol{\beta}}(D,q) is the collection of Fock space strings

{𝒴=𝒳Xq|𝒳𝜷(D),Xq𝜷(q)}.\{\mathcal{Y}=\mathcal{X}X_{q}\,|\,\mathcal{X}\in\mathcal{F}_{\boldsymbol{\beta}}(D^{*}),X_{q}\in\mathcal{F}_{\boldsymbol{\beta}}(q)\}.

For example, if X=Y=J𝜷ˇX=Y=J_{\check{\boldsymbol{\beta}\,}} (𝜷ˇ=bq+bq\check{\boldsymbol{\beta}\,}=b\cdot q+b\cdot q^{*}), then both Y(X1q)Y*(X*1_{q}) and (YX)1q(Y*X)*1_{q} are in 𝜷(q),\mathcal{F}_{\boldsymbol{\beta}}(q), but they are different, see the last example in the next subsection.

6.3. Examples of the OPE family

It is clear that 1q(q),1_{q}\in\mathcal{F}(q), so every X(D)X\in\mathcal{F}(D^{*}) belongs to the extended family. We now present some less obvious examples in the standard radial case, 𝜷ˇ=bq+bq.\check{\boldsymbol{\beta}\,}=b\cdot q+b\cdot q^{*}.

Example.

As mentioned before,

𝒪q(σ)=𝒪(σ)1q𝜷ˇ(q),\mathcal{O}_{q}^{(\sigma)}=\mathcal{O}^{(\sigma)}*1_{q}\in\mathcal{F}_{\check{\boldsymbol{\beta}\,}}(q),

and more generally 𝒪(σ)𝒪q(τ)=𝒪q(σ+τ).\mathcal{O}^{(\sigma)}*\mathcal{O}_{q}^{(\tau)}=\mathcal{O}^{(\sigma+\tau)}_{q}. We define the formal functional 𝒪q(τ,τ)\mathcal{O}_{q}^{(\tau,\tau_{*})} by

𝒪q(τ,τ):=𝒪q(τ)𝒪q(τ¯)¯=(wq)τ2/2(wq¯)τ2/2eiτΦ+(q)iτΦ(q).\mathcal{O}_{q}^{(\tau,\tau_{*})}:=\mathcal{O}_{q}^{(\tau)}\overline{\mathcal{O}_{q}^{(\overline{\tau_{*}})}}=(w_{q}^{\prime})^{\tau^{2}/2}(\overline{w_{q}^{\prime}})^{\tau_{*}^{2}/2}\mathrm{e}^{\odot i\tau\Phi^{+}(q)-i\tau_{*}\Phi^{-}(q)}.

Let us discuss the vertex algebra at the puncture. We define the multiplication of OPE exponentials at the puncture by

𝒪q(σ,σ)𝒪q(τ,τ)=𝒪(σ,σ)𝒪q(τ,τ)\mathcal{O}_{q}^{(\sigma,\sigma_{*})}\mathcal{O}_{q}^{(\tau,\tau_{*})}=\mathcal{O}^{(\sigma,\sigma_{*})}*\mathcal{O}_{q}^{(\tau,\tau_{*})}

so that 𝒪q(σ,σ)𝒪q(τ,τ)=𝒪q(σ+τ,σ+τ).\mathcal{O}_{q}^{(\sigma,\sigma_{*})}\mathcal{O}_{q}^{(\tau,\tau_{*})}=\mathcal{O}_{q}^{(\sigma+\tau,\sigma_{*}+\tau_{*})}.

Examples.

(a) We have (Φ𝜷ˇ)q=Φ(q).(\Phi_{\check{\boldsymbol{\beta}\,}})_{q}=\Phi(q). It is because (zq)w(z)/w(z)=1+o(zq).(z-q)w^{\prime}(z)/w(z)=1+o(z-q).

(b) We have

(J𝜷ˇ)q=J(q)+ib2Nw(q).(J_{\check{\boldsymbol{\beta}\,}})_{q}=J(q)+\frac{ib}{2}N_{w}(q).

This follows from J𝜷ˇ=J+j𝜷ˇJ_{\check{\boldsymbol{\beta}\,}}=J+j_{\check{\boldsymbol{\beta}\,}} with j𝜷ˇ=ib(w′′/ww/w),j_{\check{\boldsymbol{\beta}\,}}=ib({w^{\prime\prime}}/{w^{\prime}}-{w^{\prime}}/{w}), and

(ww)1q=limzq(w(z)w(z)1zq)=limzqw′′(z)w(z)+w(z)/(zq)=12w′′(q)w(q).\Big{(}\frac{w^{\prime}}{w}\Big{)}*1_{q}=\lim_{z\to q}\Big{(}\frac{w^{\prime}(z)}{w(z)}-\frac{1}{z-q}\Big{)}=\lim_{z\to q}\frac{w^{\prime\prime}(z)}{w^{\prime}(z)+w(z)/(z-q)}=\frac{1}{2}\frac{w^{\prime\prime}(q)}{w^{\prime}(q)}.

(c) We have

(T𝜷ˇ)q\displaystyle(T_{\check{\boldsymbol{\beta}\,}})_{q} =A(q)+2ibJ(q)ib2Nw(q)J(q)+(c1213b2)Sw(q)38b2Nw(q)2.\displaystyle=A(q)+2ib\partial J(q)-\frac{ib}{2}N_{w}(q)J(q)+\Big{(}\frac{c}{12}-\frac{1}{3}\,b^{2}\Big{)}S_{w}(q)-\frac{3}{8}\,b^{2}N_{w}(q)^{2}.
Example.

For Xq=X1q,X_{q}=X*1_{q}, YXq(YX)1qY*X_{q}\neq(Y*X)*1_{q} in general. For example, with the choice of X=Y=J𝜷ˇ,X=Y=J_{\check{\boldsymbol{\beta}\,}}, we have Y(X1q)(YX)1q.Y*(X*1_{q})\neq(Y*X)*1_{q}. It is because

(w′′www)21q=43Sw(q)14Nw(q)2((w′′www)1q)2=14Nw(q)2.\Big{(}\frac{w^{\prime\prime}}{w^{\prime}}-\frac{w^{\prime}}{w}\Big{)}^{2}*1_{q}=-\frac{4}{3}\,S_{w}(q)-\frac{1}{4}\,N_{w}(q)^{2}\neq\bigg{(}\Big{(}\frac{w^{\prime\prime}}{w^{\prime}}-\frac{w^{\prime}}{w}\Big{)}*1_{q}\bigg{)}^{2}=\frac{1}{4}\,N_{w}(q)^{2}.

6.4. Insertions

In this subsection we extend the insertion of the one-leg operator to the strings containing puncture functionals in particular to the fields in the extended OPE family. For simplicity, we consider the standard radial case with 𝜷ˇ=bq+bq.\check{\boldsymbol{\beta}\,}=b\cdot q+b\cdot q^{*}. Let Λζ\Lambda_{\zeta} denote the one-leg operator,

Λζ=𝒪𝜷ˇ[aζa2qa2q]=𝒪(a)(ζ)𝒪(a/2)(q)𝒪(a/2)(q)¯.\Lambda_{\zeta}=\mathcal{O}_{\check{\boldsymbol{\beta}\,}}[a\cdot\zeta-\frac{a}{2}\cdot q-\frac{a}{2}\cdot q^{*}]=\mathcal{O}^{(a)}(\zeta)\mathcal{O}^{(-a/2)}(q)\overline{\mathcal{O}^{(-a/2)}(q)}.

If qq is a node of 𝒳,\mathcal{X}, e.g.,

𝒳=𝒴Xq,Xq=X1q𝜷(q),\mathcal{X}=\mathcal{Y}X_{q},\qquad X_{q}=X*1_{q}\in\mathcal{F}_{\boldsymbol{\beta}}(q),

then the accurate definition of the product Λζ𝒳\Lambda_{\zeta}\mathcal{X} is

Λζ𝒳:=𝒪(a)(ζ)𝒴(X𝒪q),𝒪q:=𝒪(a/2)(q)𝒪(a/2)(q)¯.\Lambda_{\zeta}\mathcal{X}:=\mathcal{O}^{(a)}(\zeta)\mathcal{Y}(X*\mathcal{O}_{q}),\qquad\mathcal{O}_{q}:=\mathcal{O}^{(-a/2)}(q)\overline{\mathcal{O}^{(-a/2)}(q)}.
Proposition 6.3.

Let 𝒳=𝒴Xq,\mathcal{X}=\mathcal{Y}X_{q}, Xq=X1qX_{q}=X*1_{q} and X^q=X^1q.\widehat{X}_{q}=\widehat{X}*1_{q}. Then we have

𝐄^𝒴Xq=𝐄𝒴^X^q.\widehat{\mathbf{E}}\,\mathcal{Y}X_{q}=\mathbf{E}\,\widehat{\mathcal{Y}}\widehat{X}_{q}.
Proof.

By definition,

𝐄^𝒴Xq=𝐄Λp𝒴Xq𝐄Λp=𝐄𝒪(a)(p)𝒴(X𝒪q)𝐄Λp.\widehat{\mathbf{E}}\,\mathcal{Y}X_{q}=\frac{\mathbf{E}\,\Lambda_{p}\mathcal{Y}X_{q}}{\mathbf{E}\,\Lambda_{p}}=\frac{\mathbf{E}\,\mathcal{O}^{(a)}(p)\mathcal{Y}(X*\mathcal{O}_{q})}{\mathbf{E}\,\Lambda_{p}}.

In follows from (5.23) that

𝐄𝒴X(z)Λp𝐄Λp=𝐄𝒴^X^(z)1q.\frac{\mathbf{E}\,\mathcal{Y}X(z)\Lambda_{p}}{\mathbf{E}\,\Lambda_{p}}=\mathbf{E}\,\widehat{\mathcal{Y}}\widehat{X}(z)1_{q}.

Comparing the OPE coefficients at the puncture, we have

𝐄𝒪(a)(p)𝒴(X𝒪q)𝐄Λp=𝐄𝒴^X^q.\frac{\mathbf{E}\,\mathcal{O}^{(a)}(p)\mathcal{Y}(X*\mathcal{O}_{q})}{\mathbf{E}\,\Lambda_{p}}=\mathbf{E}\,\widehat{\mathcal{Y}}\widehat{X}_{q}.

Examples.

Let 𝜷ˇ=bq+bq\check{\boldsymbol{\beta}\,}=b\cdot q+b\cdot q^{*} and 𝜷=ap+(ba/2)q+(ba/2)q.\boldsymbol{\beta}=a\cdot p+(b-a/2)\cdot q+(b-a/2)\cdot q^{*}.

(a) Recall that (see (5.21) and Subsection 6.3, Example (a))

Φ^𝜷ˇ=Φ𝜷ˇ+aarg(1w)2w=Φ𝜷,\widehat{\Phi}_{\check{\boldsymbol{\beta}\,}}=\Phi_{\check{\boldsymbol{\beta}\,}}+a\arg\frac{(1-w)^{2}}{w}=\Phi_{\boldsymbol{\beta}},

and (Φ𝜷ˇ)q=Φ(q).(\Phi_{\check{\boldsymbol{\beta}\,}})_{q}=\Phi(q). Since (argw)1q=argwq,(\arg w)*1_{q}=\arg w_{q}^{\prime}, we have

(Φ^𝜷ˇ)q=Φ(q)aargwq.(\widehat{\Phi}_{\check{\boldsymbol{\beta}\,}})_{q}=\Phi(q)-a\arg w_{q}^{\prime}.

(b) It follows from (5.24) and Example (b) in Subsection 6.3 that

(J^𝜷ˇ)q=J(q)+i(a4+b2)Nw(q)+iaw(q).(\widehat{J}_{\check{\boldsymbol{\beta}\,}})_{q}=J(q)+i\Big{(}\frac{a}{4}+\frac{b}{2}\Big{)}N_{w}(q)+iaw^{\prime}(q).

(c) Note 𝒪^q(σ):=𝒪^(σ)1q=(wq)aσ/2𝒪q(σ),\widehat{\mathcal{O}}_{q}^{(\sigma)}:=\widehat{\mathcal{O}}^{(\sigma)}*1_{q}=(w_{q}^{\prime})^{-a\sigma/2}\mathcal{O}_{q}^{(\sigma)}, cf.

𝒪^(σ)=(1ww)aσ𝒪(σ).\widehat{\mathcal{O}}^{(\sigma)}=\Big{(}\frac{1-w}{\sqrt{w}}\Big{)}^{a\sigma}\mathcal{O}^{(\sigma)}.

We have 𝐄^𝒪q(σ)=𝐄𝒪^q(σ)\widehat{\mathbf{E}}\,\mathcal{O}_{q}^{(\sigma)}=\mathbf{E}\,\widehat{\mathcal{O}}_{q}^{(\sigma)} because the differentials on both sides have the same dimension and are equal to 1 in (𝔻,0,1).(\mathbb{D},0,1).

It is not difficult to see that Theorem 5.5 extends to Theorem 1.1 using the argument in the proof of Proposition 6.3. We remark that Proposition 6.3 is a special case (𝜷ˇ=bq+bq\check{\boldsymbol{\beta}\,}=b\cdot q+b\cdot q^{*} and 𝜷=ap+(ba/2)q+(ba/2)q\boldsymbol{\beta}=a\cdot p+(b-a/2)\cdot q+(b-a/2)\cdot q^{*}) of Theorem 1.1.

For two background charges 𝜷,𝜷ˇ\boldsymbol{\beta},\check{\boldsymbol{\beta}\,} satisfying the neutrality condition (NCb),(\mathrm{NC}_{b}), let

𝐄^𝒳𝜷ˇ:=𝐄𝒳𝜷ˇV[𝜷𝜷ˇ]=𝐄𝒳𝜷ˇ𝒪𝜷ˇ[𝜷𝜷ˇ]𝐄𝒪𝜷ˇ[𝜷𝜷ˇ].\widehat{\mathbf{E}}\,\mathcal{X}_{\check{\boldsymbol{\beta}\,}}:=\mathbf{E}\,\mathcal{X}_{\check{\boldsymbol{\beta}\,}}V^{\odot}[\boldsymbol{\beta}-\check{\boldsymbol{\beta}\,}]=\frac{\mathbf{E}\,\mathcal{X}_{\check{\boldsymbol{\beta}\,}}\mathcal{O}_{\check{\boldsymbol{\beta}\,}}[\boldsymbol{\beta}-\check{\boldsymbol{\beta}\,}]}{\mathbf{E}\,\mathcal{O}_{\check{\boldsymbol{\beta}\,}}[\boldsymbol{\beta}-\check{\boldsymbol{\beta}\,}]}.

If the nodes of 𝒳𝜷ˇ\mathcal{X}_{\check{\boldsymbol{\beta}\,}} intersect with the those of 𝜷𝜷ˇ\boldsymbol{\beta}-\check{\boldsymbol{\beta}\,}, then the product of 𝒳𝜷ˇ\mathcal{X}_{\check{\boldsymbol{\beta}\,}} and 𝒪𝜷ˇ[𝜷𝜷ˇ]\mathcal{O}_{\check{\boldsymbol{\beta}\,}}[\boldsymbol{\beta}-\check{\boldsymbol{\beta}\,}] should be understood as the sense of OPE’s.

The insertion of V[𝜷𝜷ˇ]V^{\odot}[\boldsymbol{\beta}-\check{\boldsymbol{\beta}\,}] is an operator 𝒳𝒳^\mathcal{X}\mapsto\widehat{\mathcal{X}} on Fock space functionals/fields given by the rules (5.20) and the formula

Φ^𝜷ˇ=Φ𝜷.\widehat{\Phi}_{\check{\boldsymbol{\beta}\,}}=\Phi_{\boldsymbol{\beta}}.
Example.

Suppose that a divisor 𝝉\boldsymbol{\tau} on SS satisfies the neutrality condition (NC0).(\mathrm{NC}_{0}). Then we have

(6.2) 𝒪^𝜷ˇ[𝝉]=𝒪𝜷[𝝉].\widehat{\mathcal{O}}_{\check{\boldsymbol{\beta}\,}}[\boldsymbol{\tau}]=\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}].

Since both sides contain the same Wick’s exponential V[𝝉]=eiΦ[𝝉],V^{\odot}[\boldsymbol{\tau}]=\mathrm{e}^{\odot i\Phi[\boldsymbol{\tau}]}, all we need is to check that both sides have the same correlation functions. It follows from Theorem 1.1 and the algebra of OPE exponentials that

𝐄𝒪^𝜷ˇ[𝝉]=𝐄^𝒪𝜷ˇ[𝝉]=𝐄𝒪𝜷ˇ[𝝉]𝒪𝜷ˇ[𝜷𝜷ˇ]𝐄𝒪𝜷ˇ[𝜷𝜷ˇ]=𝐄𝒪𝜷ˇ[𝝉+𝜷𝜷ˇ]𝐄𝒪𝜷ˇ[𝜷𝜷ˇ]=C[𝜷+𝝉]C[𝜷]=𝐄𝒪𝜷[𝝉].\mathbf{E}\,\widehat{\mathcal{O}}_{\check{\boldsymbol{\beta}\,}}[\boldsymbol{\tau}]=\widehat{\mathbf{E}}\,\mathcal{O}_{\check{\boldsymbol{\beta}\,}}[\boldsymbol{\tau}]=\frac{\mathbf{E}\,\mathcal{O}_{\check{\boldsymbol{\beta}\,}}[\boldsymbol{\tau}]\mathcal{O}_{\check{\boldsymbol{\beta}\,}}[\boldsymbol{\beta}-\check{\boldsymbol{\beta}\,}]}{\mathbf{E}\,\mathcal{O}_{\check{\boldsymbol{\beta}\,}}[\boldsymbol{\beta}-\check{\boldsymbol{\beta}\,}]}=\frac{\mathbf{E}\,\mathcal{O}_{\check{\boldsymbol{\beta}\,}}[\boldsymbol{\tau}+\boldsymbol{\beta}-\check{\boldsymbol{\beta}\,}]}{\mathbf{E}\,\mathcal{O}_{\check{\boldsymbol{\beta}\,}}[\boldsymbol{\beta}-\check{\boldsymbol{\beta}\,}]}=\frac{C[\boldsymbol{\beta}+\boldsymbol{\tau}]}{C[\boldsymbol{\beta}]}=\mathbf{E}\,\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}].

7. Ward identities and BPZ equations

We represent the Ward functionals as the Lie derivative operators within correlations of fields in the extended OPE family. Combining these representations (Ward identities) with the expression of the Virasoro fields in terms of the Ward functionals, we derive the Belavin-Polyakov-Zamolodchikov equations (BPZ equations) for correlations involving the Virasoro fields or the Virasoro generators.

7.1. Ward’s identity

Let D:=D(supp𝜷+supp𝜷)D^{*}:=D\setminus(\mathrm{supp}\,\boldsymbol{\beta}^{+}\cup\,\mathrm{supp}\,\boldsymbol{\beta}^{-}) and 𝜷=𝜷++𝜷\boldsymbol{\beta}=\boldsymbol{\beta}^{+}+\boldsymbol{\beta}^{-}_{*} as before.

Lemma 7.1.

For a non-random local vector field v,v, all fields in 𝛃(D)\mathcal{F}_{\boldsymbol{\beta}}(D^{*}) satisfy local Ward’s identity (5.4) in Dhol(v)D.D_{\mathrm{hol}}(v)\cap D^{*}.

Proof.

For simplicity, let us consider the case of holomorphic fields XX and YY only. We need to show that if XX and YY satisfy the residue form (5.4) of Ward’s identity at the nodes, then so does Xα+nY𝜷(D).X*_{\alpha+n}Y\in\mathcal{F}_{\boldsymbol{\beta}}(D^{*}). (The case α=0\alpha=0 was covered in [18, Proposition 5.8].) In this case, we have v+X=(vT)1X\mathcal{L}_{v}^{+}X=(vT)*_{-1}X and v+Y=(vT)1Y.\mathcal{L}_{v}^{+}Y=(vT)*_{-1}Y. We need to check that

v+(Xα+nY)=(vT)1(Xα+nY).\mathcal{L}_{v}^{+}(X*_{\alpha+n}Y)=(vT)*_{-1}(X*_{\alpha+n}Y).

By Leibniz’s rule, the left-hand side is (v+X)α+nY+Xα+n(v+Y).(\mathcal{L}_{v}^{+}X)*_{\alpha+n}Y+X*_{\alpha+n}(\mathcal{L}_{v}^{+}Y). Since XX and YY satisfy the residue form of Ward’s identity, it is equal to

[(vT)1X]α+nY+Xα+n[(vT)1Y].[(vT)*_{-1}X]*_{\alpha+n}Y+X*_{\alpha+n}[(vT)*_{-1}Y].

Let us show [(vT)1X]α+nY+Xα+n[(vT)1Y]=(vT)1(Xα+nY).[(vT)*_{-1}X]*_{\alpha+n}Y+X*_{\alpha+n}[(vT)*_{-1}Y]=(vT)*_{-1}(X*_{\alpha+n}Y).

Let C,C_{-}, C,C, and C+C_{+} be three concentric circles centered at zz with increasing radii, ε<ε<ε+(1).\varepsilon_{-}<\varepsilon<\varepsilon_{+}(\ll 1). In the correlations with the string 𝒳\mathcal{X} whose nodes are outside of the discs, we have

(vT)1(Xα+nY)(z)=12πiζC+v(ζ)T(ζ)(Xα+nY)(z)d\displaystyle(vT)*_{-1}(X*_{\alpha+n}Y)(z)=\frac{1}{2\pi i}\,\oint_{\zeta\in C_{+}}v(\zeta)T(\zeta)\,(X*_{\alpha+n}Y)(z)\,\mathrm{d} ζ\displaystyle\zeta
=1(2πi)2ζC+ηCv(ζ)T(ζ)(ηz)αn1X(η)Y(z)dηd\displaystyle=\frac{1}{(2\pi i)^{2}}\,\oint_{\zeta\in C_{+}}\oint_{\eta\in C}v(\zeta)T(\zeta)\,(\eta-z)^{-\alpha-n-1}X(\eta)\,Y(z)\,\mathrm{d}\eta\,\mathrm{d} ζ.\displaystyle\zeta.

For α,\alpha\notin\mathbb{Z}, the multivalued factor (ηz)α(\eta-z)^{-\alpha} is canceled out by the OPEα\mathrm{OPE}_{\alpha} of X(η)X(\eta) and Y(z).Y(z). In a similar way, we compute Xα+n[(vT)1Y](z)X*_{\alpha+n}[(vT)*_{-1}Y](z) as

1(2πi)2ζCηCv(ζ)T(ζ)(ηz)αn1X(η)Y(z)dηdζ.\qquad\frac{1}{(2\pi i)^{2}}\,\oint_{\zeta\in C_{-}}\oint_{\eta\in C}v(\zeta)T(\zeta)\,(\eta-z)^{-\alpha-n-1}X(\eta)\,Y(z)\,\mathrm{d}\eta\,\mathrm{d}\zeta.

Subtracting, we express (vT)1(Xα+nY)(z)Xα+n[(vT)1Y](z)(vT)*_{-1}(X*_{\alpha+n}Y)(z)-X*_{\alpha+n}[(vT)*_{-1}Y](z) as

1(2πi)2ηCζ[C+C]v(ζ)T(ζ)(ηz)αn1X(η)Y(z)dζdη\displaystyle\frac{1}{(2\pi i)^{2}}\,\oint_{\eta\in C}\oint_{\zeta\in[C_{+}-C_{-}]}v(\zeta)T(\zeta)\,(\eta-z)^{-\alpha-n-1}X(\eta)\,Y(z)\,\mathrm{d}\zeta\,\mathrm{d}\eta
=1(2πi)2ηC(ηz)αn1(η)v(ζ)T(ζ)X(η)dζY(z)dη.\displaystyle=\frac{1}{(2\pi i)^{2}}\,\oint_{\eta\in C}(\eta-z)^{-\alpha-n-1}\oint_{(\eta)}v(\zeta)T(\zeta)X(\eta)\,\mathrm{d}\zeta\,Y(z)\,\mathrm{d}\eta.

The last integral simplifies to [(vT)1X]α+nY(z).[(vT)*_{-1}X]*_{\alpha+n}Y(z).

The next lemma states that within correlations of fields in 𝜷(D)\mathcal{F}_{\boldsymbol{\beta}}(D^{*}) the residue form of Ward’s identity holds at the marked points qkq_{k}’s where the background charges are placed.

Lemma 7.2.

We have

12πi(qk)vA𝜷=v+(qk)=v(qk)qk\frac{1}{2\pi i}\oint_{(q_{k})}vA_{\boldsymbol{\beta}}=\mathcal{L}_{v}^{+}(q_{k})=v(q_{k})\partial_{q_{k}}

within correlations of fields in 𝛃(D).\mathcal{F}_{\boldsymbol{\beta}}(D^{*}).

Proof.

From the relation

A𝜷=A+ibJj𝜷J,j𝜷(ζ)=𝐄J𝜷(ζ),A_{\boldsymbol{\beta}}=A+ib\partial J-j_{\boldsymbol{\beta}}J,\qquad j_{\boldsymbol{\beta}}(\zeta)=\mathbf{E}\,J_{\boldsymbol{\beta}}(\zeta),

and the fact that the meromorphic function j𝜷j_{\boldsymbol{\beta}} has a simple pole at qkq_{k} with Resqkj𝜷=iβk,\operatorname{Res}_{q_{k}}j_{\boldsymbol{\beta}}=-i\beta_{k}, it follows that

12πi(qk)vA𝜷=12πi(qk)vj𝜷J=iβkv(qk)J(qk).\frac{1}{2\pi i}\oint_{(q_{k})}vA_{\boldsymbol{\beta}}=-\frac{1}{2\pi i}\oint_{(q_{k})}vj_{\boldsymbol{\beta}}J=i\beta_{k}v(q_{k})J(q_{k}).

We now claim that

(7.1) iβk𝐄J(qk)𝒳𝜷=qk𝐄𝒳𝜷i\beta_{k}\mathbf{E}\,J(q_{k})\mathcal{X}_{\boldsymbol{\beta}}=\partial_{q_{k}}\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}}

for any string 𝒳𝜷\mathcal{X}_{\boldsymbol{\beta}} of fields in 𝜷(D).\mathcal{F}_{\boldsymbol{\beta}}(D^{*}). Then lemma follows immediately. Note that the second identity

v+(qk)𝐄𝒳𝜷=v(qk)qk𝐄𝒳𝜷\mathcal{L}_{v}^{+}(q_{k})\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}}=v(q_{k})\partial_{q_{k}}\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}}

is obvious since 𝐄𝒳𝜷\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}} is a scalar with respect to qkq_{k}.

To prove the claim, let us consider a reference background charge 𝜷0\boldsymbol{\beta}_{0} such that

supp𝜷supp𝜷0=.\mathrm{supp}\,\boldsymbol{\beta}\cap\mathrm{supp}\,\boldsymbol{\beta}_{0}=\emptyset.

Differentiating the relation (Theorem 5.5) 𝐄𝒳𝜷=𝐄𝒳𝜷0eiΦ[𝜷𝜷0],\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}}=\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}_{0}}\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]}, we have

(7.2) qk𝐄𝒳𝜷=qk𝐄𝒳𝜷0eiΦ[𝜷𝜷0]=iβk𝐄𝒳𝜷0J(qk)eiΦ[𝜷𝜷0].\partial_{q_{k}}\mathbf{E}\mathcal{X}_{\boldsymbol{\beta}}=\partial_{q_{k}}\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}_{0}}\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]}=i\beta_{k}\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}_{0}}J(q_{k})\odot\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]}.

It follows from Wick’s calculus that

(7.3) J(qk)eiΦ[𝜷𝜷0]=J(qk)eiΦ[𝜷𝜷0]i𝐄[J(qk)Φ[𝜷𝜷0]]eiΦ[𝜷𝜷0].J(q_{k}^{\prime})\odot\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]}=J(q_{k}^{\prime})\,\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]}-i\mathbf{E}[J(q_{k}^{\prime})\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]]\,\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]}.

(Indeed, if one applies Wick’s calculus to J(qk)eiΦ[𝜷𝜷0],J(q_{k}^{\prime})\,\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]}, the term J(qk)eiΦ[𝜷𝜷0]J(q_{k}^{\prime})\odot\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]} comes from 0 contraction and the term i𝐄[J(qk)Φ[𝜷𝜷0]]eiΦ[𝜷𝜷0]i\mathbf{E}[J(q_{k}^{\prime})\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]]\,\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]} comes from 11 contraction.) Here we take qkqkq_{k}^{\prime}\neq q_{k} so that the term J𝜷0(qk)eiΦ[𝜷𝜷0]J_{\boldsymbol{\beta}_{0}}(q_{k}^{\prime})\,\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]} in the above makes sense but send qkq_{k}^{\prime} to qkq_{k} at the end. By Wick’s calculus, we find the non-random factor in the last term as

(7.4) i𝐄[J(qk)Φ[𝜷𝜷0]]=𝐄[J(qk)eiΦ[𝜷𝜷0]]=j𝜷(qk)j𝜷0(qk).i\mathbf{E}[J(q_{k}^{\prime})\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]]=\mathbf{E}[J(q_{k}^{\prime})\,\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]}]=j_{\boldsymbol{\beta}}(q_{k}^{\prime})-j_{\boldsymbol{\beta}_{0}}(q_{k}^{\prime}).

Alternatively, the last identity in the above follows from Theorem 5.5:

j𝜷(qk)=𝐄J𝜷(qk)\displaystyle j_{\boldsymbol{\beta}}(q_{k}^{\prime})=\mathbf{E}\,J_{\boldsymbol{\beta}}(q_{k}^{\prime}) =𝐄[J𝜷0(qk)eiΦ[𝜷𝜷0]]\displaystyle=\mathbf{E}[J_{\boldsymbol{\beta}_{0}}(q_{k}^{\prime})\,\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]}]
=𝐄[(J(qk)+j𝜷0(qk))eiΦ[𝜷𝜷0]]=𝐄[J(qk)eiΦ[𝜷𝜷0]]+j𝜷0(qk).\displaystyle=\mathbf{E}[(J(q_{k}^{\prime})+j_{\boldsymbol{\beta}_{0}}(q_{k}^{\prime}))\,\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]}]=\mathbf{E}[J(q_{k}^{\prime})\,\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]}]+j_{\boldsymbol{\beta}_{0}}(q_{k}^{\prime}).

Combining (7.3) and (7.4), we have

J(qk)eiΦ[𝜷𝜷0]=J(qk)eiΦ[𝜷𝜷0]+(j𝜷0(qk)j𝜷(qk))eiΦ[𝜷𝜷0].J(q_{k}^{\prime})\odot\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]}=J(q_{k}^{\prime})\,\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]}+\big{(}j_{\boldsymbol{\beta}_{0}}(q_{k}^{\prime})-j_{\boldsymbol{\beta}}(q_{k}^{\prime})\big{)}\,\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]}.

In application to 𝒳𝜷0\mathcal{X}_{\boldsymbol{\beta}_{0}} within correlations,

𝐄𝒳𝜷0J(qk)eiΦ[𝜷𝜷0]=𝐄𝒳𝜷0J𝜷0(qk)eiΦ[𝜷𝜷0]j𝜷(qk)𝐄𝒳𝜷0eiΦ[𝜷𝜷0].\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}_{0}}J(q_{k}^{\prime})\odot\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]}=\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}_{0}}J_{\boldsymbol{\beta}_{0}}(q_{k}^{\prime})\,\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]}-j_{\boldsymbol{\beta}}(q_{k}^{\prime})\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}_{0}}\,\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]}.

Applying Theorem 5.5 again to each term on the right-hand side, we obtain

𝐄𝒳𝜷0J(qk)eiΦ[𝜷𝜷0]=𝐄𝒳𝜷J𝜷(qk)j𝜷(qk)𝐄𝒳𝜷=𝐄J(qk)𝒳𝜷.\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}_{0}}J(q_{k}^{\prime})\odot\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]}=\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}}J_{\boldsymbol{\beta}}(q_{k}^{\prime})-j_{\boldsymbol{\beta}}(q_{k}^{\prime})\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}}=\mathbf{E}\,J(q_{k}^{\prime})\mathcal{X}_{\boldsymbol{\beta}}.

Sending qkq_{k}^{\prime} to qk,q_{k}, we have

𝐄𝒳𝜷0J(qk)eiΦ[𝜷𝜷0]=𝐄J(qk)𝒳𝜷.\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}_{0}}J(q_{k})\odot\mathrm{e}^{\odot i\Phi[\boldsymbol{\beta}-\boldsymbol{\beta}_{0}]}=\mathbf{E}\,J(q_{k})\mathcal{X}_{\boldsymbol{\beta}}.

The claim (7.1) now follows from the above equation and (7.2). We now finish the proof of the lemma. ∎

Combining the above two lemmas, we claim that all fields in 𝜷(D)\mathcal{F}_{\boldsymbol{\beta}}(D^{*}) satisfy Ward’s identity (5.4) in Dhol(v)DD_{\mathrm{hol}}(v)\cap D within correlations. Applying the same arguments as in the proof of Lemma 7.1 to this claim, we extend this claim to 𝜷(D).\mathcal{F}_{\boldsymbol{\beta}}(D).

Theorem 7.3.

For a non-random local vector field v,v, all fields in 𝛃(D)\mathcal{F}_{\boldsymbol{\beta}}(D) satisfy local Ward’s identity (5.4) in Dhol(v)DD_{\mathrm{hol}}(v)\cap D within correlations.

7.2. Ward equations on the Riemann sphere

In this subsection we express the Virasoro fields T𝜷T_{\boldsymbol{\beta}} in terms of Lie derivative within correlations of fields in 𝜷(S).\mathcal{F}_{\boldsymbol{\beta}}(S).

Given a meromorphic vector field with poles ξl,\xi_{l}, we define the Ward functional W𝜷+(v)W_{\boldsymbol{\beta}}^{+}(v) by

𝐄W𝜷+(v)𝒳:=12πil(ξl)v𝐄A𝜷𝒳\mathbf{E}\,W_{\boldsymbol{\beta}}^{+}(v)\mathcal{X}:=-\frac{1}{2\pi i}\sum_{l}\oint_{(\xi_{l})}v\,\mathbf{E}\,A_{\boldsymbol{\beta}}\mathcal{X}

for any tensor product 𝒳\mathcal{X} of Fock space fields such that the set S𝒳S_{\mathcal{X}} of all nodes of 𝒳\mathcal{X} does not intersect the set of poles of v.v.

Proposition 7.4.

In correlations with any string of fields in the extended OPE family 𝛃(S),\mathcal{F}_{\boldsymbol{\beta}}(S), we have

W𝜷+(v)=v+.W_{\boldsymbol{\beta}}^{+}(v)=\mathcal{L}_{v}^{+}.
Proof.

Since the sum of all residues of the meromorphic 11-differential ζv(ζ)𝐄A𝜷(ζ)𝒳𝜷\zeta\mapsto v(\zeta)\mathbf{E}\,A_{\boldsymbol{\beta}}(\zeta)\mathcal{X}_{\boldsymbol{\beta}} is zero, we have

𝐄W𝜷+(v)𝒳𝜷=12πij(zj)v𝐄A𝜷𝒳𝜷+12πik(qk)v𝐄A𝜷𝒳𝜷.\mathbf{E}\,W_{\boldsymbol{\beta}}^{+}(v)\mathcal{X}_{\boldsymbol{\beta}}=\frac{1}{2\pi i}\sum_{j}\oint_{(z_{j})}v\,\mathbf{E}\,A_{\boldsymbol{\beta}}\mathcal{X}_{\boldsymbol{\beta}}+\frac{1}{2\pi i}\sum_{k}\oint_{(q_{k})}v\,\mathbf{E}\,A_{\boldsymbol{\beta}}\mathcal{X}_{\boldsymbol{\beta}}.

It follows from Ward’s identity (Theorem 7.3) that

12πij(zj)v𝐄A𝜷𝒳𝜷+12πik(qk)v𝐄A𝜷𝒳𝜷=j𝐄v+(zj)𝒳𝜷+k𝐄v+(qk)𝒳𝜷=𝐄v+𝒳𝜷,\frac{1}{2\pi i}\sum_{j}\oint_{(z_{j})}v\,\mathbf{E}\,A_{\boldsymbol{\beta}}\mathcal{X}_{\boldsymbol{\beta}}+\frac{1}{2\pi i}\sum_{k}\oint_{(q_{k})}v\,\mathbf{E}\,A_{\boldsymbol{\beta}}\mathcal{X}_{\boldsymbol{\beta}}=\sum_{j}\mathbf{E}\,\mathcal{L}_{v}^{+}(z_{j})\mathcal{X}_{\boldsymbol{\beta}}+\sum_{k}\mathbf{E}\,\mathcal{L}_{v}^{+}(q_{k})\mathcal{X}_{\boldsymbol{\beta}}=\mathbf{E}\,\mathcal{L}_{v}^{+}\mathcal{X}_{\boldsymbol{\beta}},

which completes the proof. ∎

Given ξ,\xi\in\mathbb{C}, let us consider the vector fields kξ,vξk_{\xi},v_{\xi} given by

kξ(z)=1ξz,vξ(z)=zξ+zξzk_{\xi}(z)=\frac{1}{\xi-z},\qquad v_{\xi}(z)=z\frac{\xi+z}{\xi-z}

in the identity chart of .\mathbb{C}. Ward’s equations on the Riemann sphere now follow from the previous theorem.

Corollary 7.5.

In the ^\widehat{\mathbb{C}}-uniformization, for any tensor product 𝒳𝛃\mathcal{X}_{\boldsymbol{\beta}} of fields in the extended OPE family 𝛃,\mathcal{F}_{\boldsymbol{\beta}}, we have

(7.5) 𝐄T𝜷(ξ)𝒳𝜷=𝐄T𝜷(ξ)𝐄𝒳𝜷+𝐄kξ+𝒳𝜷\mathbf{E}\,T_{\boldsymbol{\beta}}(\xi)\mathcal{X}_{\boldsymbol{\beta}}=\mathbf{E}\,T_{\boldsymbol{\beta}}(\xi)\,\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}}+\mathbf{E}\,\mathcal{L}_{k_{\xi}}^{+}\mathcal{X}_{\boldsymbol{\beta}}

and

(7.6) 2ξ2𝐄T𝜷(ξ)𝒳𝜷=2ξ2𝐄T𝜷(ξ)𝐄𝒳𝜷+𝐄vξ+𝒳𝜷.2\xi^{2}\mathbf{E}\,T_{\boldsymbol{\beta}}(\xi)\mathcal{X}_{\boldsymbol{\beta}}=2\xi^{2}\mathbf{E}\,T_{\boldsymbol{\beta}}(\xi)\,\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}}+\mathbf{E}\,\mathcal{L}_{v_{\xi}}^{+}\mathcal{X}_{\boldsymbol{\beta}}.

7.3. Ward’s equations in a simply-connected domain

To derive Ward’s equations in \mathbb{H} and 𝔻\mathbb{D}, we modify Ward’s functional to be defined in a simply-connected domain. Ward’s equations in \mathbb{H} (resp. in 𝔻\mathbb{D}) represent the insertion of the Virasoro fields within correlations of fields in 𝜷\mathcal{F}_{\boldsymbol{\beta}} in terms of Lie derivative operators with respect to the chordal (resp. radial) Loewner vector fields.

We first consider the special case that no background charge is placed on the boundary D.\partial D. The general case can be treated through a suitable limit procedure. In this special case, A𝜷A_{\boldsymbol{\beta}} is continuous on the boundary. (The continuity on the boundary should be understood in terms of standard boundary charts.) As we mentioned before, we mostly concern ourselves with a symmetric background charge 𝜷.\boldsymbol{\beta}. In this case, A𝜷A_{\boldsymbol{\beta}} is real on the boundary. (Again, we understand the real-valuedness on the boundary in terms of standard boundary charts.)

Given a meromorphic vector field vv in DD (continuous up to the boundary) with poles ξl\xi_{l}’s (ξlD\xi_{l}\in D), we define the Ward functional W𝜷+(v)W_{\boldsymbol{\beta}}^{+}(v) by

W𝜷+(v):=limε0(12πiDεvA𝜷1πDε(¯v)A𝜷),W_{\boldsymbol{\beta}}^{+}(v):=\lim_{\varepsilon\to 0}\Big{(}\frac{1}{2\pi i}\int_{\partial D_{\varepsilon}}vA_{\boldsymbol{\beta}}-\frac{1}{\pi}\int_{D_{\varepsilon}}({\bar{\partial}}v)A_{\boldsymbol{\beta}}\Big{)},

where Dε=DB(ξl,ε).D_{\varepsilon}=D\setminus\bigcup B(\xi_{l},\varepsilon). Both integrals are coordinate independent since vA𝜷vA_{\boldsymbol{\beta}} is a [1,0][1,0]-differential and (¯v)A𝜷({\bar{\partial}}v)A_{\boldsymbol{\beta}} is a [1,1][1,1]-differential. Their correlations with Fock space functionals 𝒳\mathcal{X} are well-defined provided that any node of 𝒳\mathcal{X} is not a pole of v.v.

Somewhat symbolically, the Ward functional W𝜷+(v)W_{\boldsymbol{\beta}}^{+}(v) can be represented as

W𝜷+(v)=12πiDvA𝜷l12πi(ξl)vA𝜷,W_{\boldsymbol{\beta}}^{+}(v)=\frac{1}{2\pi i}\int_{\partial D}vA_{\boldsymbol{\beta}}-\sum_{l}\frac{1}{2\pi i}\oint_{(\xi_{l})}vA_{\boldsymbol{\beta}},

or

W𝜷+(v)=12πiDvA𝜷1πD(¯v)A𝜷,W_{\boldsymbol{\beta}}^{+}(v)=\frac{1}{2\pi i}\int_{\partial D}vA_{\boldsymbol{\beta}}-\frac{1}{\pi}\int_{D}({\bar{\partial}}v)A_{\boldsymbol{\beta}},

where ¯v{\bar{\partial}}v in the last integral should be understood in the sense of distributions. If a background charge is placed on the boundary, then the first integral should be taken in the sense of the Cauchy principal value.

From now on, we only consider the case that 𝜷\boldsymbol{\beta} is symmetric. Then W𝜷=(A𝜷,A𝜷¯).W_{\boldsymbol{\beta}}=(A_{\boldsymbol{\beta}},\overline{A_{\boldsymbol{\beta}}}). We now prove Theorem 1.2.

Proof of Theorem 1.2.

We may consider the case that supp𝜷±D.\mathrm{supp}\,\boldsymbol{\beta}^{\pm}\subseteq D. It is enough to show that

W𝜷+(v)=v+.W_{\boldsymbol{\beta}}^{+}(v)=\mathcal{L}_{v}^{+}.

in correlations with any string of fields in the extended OPE family 𝜷(D).\mathcal{F}_{\boldsymbol{\beta}}(D).

Let Dj(Dk,Dl′′)D_{j}\,(D^{\prime}_{k},D^{\prime\prime}_{l}) be a sufficiently small open disk centered at zj(qk,ξl),z_{j}\,(q_{k},\xi_{l}), respectively. Applying Stokes’ theorem to

d(v(z)𝐄A𝜷(z)𝒳𝜷dz)=¯(v(z)𝐄A𝜷(z)𝒳𝜷)dzdz¯=2i¯(v(z)𝐄A𝜷(z)𝒳𝜷)dxdy\mathrm{d}\big{(}v(z)\mathbf{E}\,A_{\boldsymbol{\beta}}(z)\mathcal{X}_{\boldsymbol{\beta}}\,\mathrm{d}z\big{)}=-{\bar{\partial}}\big{(}v(z)\mathbf{E}\,A_{\boldsymbol{\beta}}(z)\mathcal{X}_{\boldsymbol{\beta}}\big{)}\,\mathrm{d}z\wedge\mathrm{d}\bar{z}=2i{\bar{\partial}}\big{(}v(z)\mathbf{E}\,A_{\boldsymbol{\beta}}(z)\mathcal{X}_{\boldsymbol{\beta}}\big{)}\,\mathrm{d}x\wedge\mathrm{d}y

over D(jDjkDklDl′′),D\setminus(\bigcup_{j}D_{j}\cup\bigcup_{k}D^{\prime}_{k}\cup\bigcup_{l}D^{\prime\prime}_{l}), we have

𝐄W𝜷+(v)𝒳𝜷=12πij(zj)v𝐄A𝜷𝒳𝜷+12πij(qk)v𝐄A𝜷𝒳𝜷.\mathbf{E}\,W_{\boldsymbol{\beta}}^{+}(v)\mathcal{X}_{\boldsymbol{\beta}}=\frac{1}{2\pi i}\sum_{j}\oint_{(z_{j})}v\,\mathbf{E}\,A_{\boldsymbol{\beta}}\mathcal{X}_{\boldsymbol{\beta}}+\frac{1}{2\pi i}\sum_{j}\oint_{(q_{k})}v\,\mathbf{E}\,A_{\boldsymbol{\beta}}\mathcal{X}_{\boldsymbol{\beta}}.

Theorem now follows from Ward’s identity (Theorem 7.3) as in the proof of Proposition 7.4. ∎

We recall the representation of a stress tensor AA in terms of Ward’s functionals W𝜷+(v)W_{\boldsymbol{\beta}}^{+}(v) with the meromorphic vector field v=kζ(ζ).v=k_{\zeta}(\zeta\in\mathbb{C}).

Proposition 7.6 (Proposition 5.11 in [18]).

Let AA be a holomorphic quadratic differential in ,\mathbb{H}, and W=(A,A¯).W=(A,\bar{A}). Suppose AA is continuous and real on the boundary (including \infty). Then

(7.7) (Aid)(ξ)=W+(kξ)+W+(kξ¯)¯.(A\,\|\,\mathrm{id})(\xi)=W^{+}(k_{\xi})+\overline{W^{+}(k_{\bar{\xi}})}.

Applying the above proposition to A𝜷A_{\boldsymbol{\beta}} in the case that supp𝜷±D,\mathrm{supp}\,\boldsymbol{\beta}^{\pm}\subseteq D, we obtain (7.7) for A=A𝜷A=A_{\boldsymbol{\beta}} and W+=W𝜷+.W^{+}=W^{+}_{\boldsymbol{\beta}}. Combining the above proposition with Theorem 1.2 and using a limit procedure in the general case that supp𝜷±D¯,\mathrm{supp}\,\boldsymbol{\beta}^{\pm}\subseteq\mkern 1.5mu\overline{\mkern-1.5muD\mkern-1.5mu}\mkern 1.5mu, we derive Ward’s equation in .\mathbb{H}.

Corollary 7.7 (Ward’s equations in \mathbb{H}).

In the \mathbb{H}-uniformization, for any tensor product 𝒳𝛃\mathcal{X}_{\boldsymbol{\beta}} of fields in the extended OPE family 𝛃,\mathcal{F}_{\boldsymbol{\beta}}, we have

(7.8) 𝐄T𝜷(ξ)𝒳𝜷=𝐄T𝜷(ξ)𝐄𝒳𝜷+𝐄(kξ++kξ¯)𝒳𝜷.\mathbf{E}\,T_{\boldsymbol{\beta}}(\xi)\mathcal{X}_{\boldsymbol{\beta}}=\mathbf{E}\,T_{\boldsymbol{\beta}}(\xi)\,\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}}+\mathbf{E}\,(\mathcal{L}_{k_{\xi}}^{+}+\mathcal{L}_{k_{\bar{\xi}}}^{-})\mathcal{X}_{\boldsymbol{\beta}}.
Example.

Let 𝒳𝜷=X1(z1)Xn(zn)\mathcal{X}_{\boldsymbol{\beta}}=X_{1}(z_{1})\cdots X_{n}(z_{n}) be the tensor product of [λj+,λj][\lambda_{j}^{+},\lambda_{j}^{-}]-differentials XjX_{j} in 𝜷.\mathcal{F}_{\boldsymbol{\beta}}. If ξ,\xi\in\mathbb{R}, then we have

𝐄T𝜷(ξ)𝒳𝜷\displaystyle\mathbf{E}\,T_{\boldsymbol{\beta}}(\xi)\mathcal{X}_{\boldsymbol{\beta}} =𝐄T𝜷(ξ)𝐄𝒳𝜷+k(qkξqk+¯qkξq¯k)𝐄𝒳𝜷\displaystyle=\mathbf{E}\,T_{\boldsymbol{\beta}}(\xi)\,\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}}+\sum_{k}\Big{(}\frac{\partial_{q_{k}}}{\xi-q_{k}}+\frac{{\bar{\partial}}_{q_{k}}}{\xi-\bar{q}_{k}}\Big{)}\,\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}}
+j(jξzj+λj+(ξzj)2+¯jξz¯j+λj(ξz¯j)2)𝐄𝒳𝜷.\displaystyle+\sum_{j}\Big{(}\frac{\partial_{j}}{\xi-z_{j}}+\frac{\lambda_{j}^{+}}{(\xi-z_{j})^{2}}+\frac{{\bar{\partial}}_{j}}{\xi-\bar{z}_{j}}+\frac{\lambda_{j}^{-}}{(\xi-\bar{z}_{j})^{2}}\Big{)}\,\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}}.

Given a meromorphic vector field vv with a local flow ztz_{t} on ^,\widehat{\mathbb{C}}, we define its reflected vector field vv^{*} with respect to the unit circle 𝔻\partial\mathbb{D} by the vector field of the reflected flow ztz_{t}^{*} of zt.z_{t}. Indeed, if ztz_{t} is the flow of vv, then zt=1/z¯tz_{t}^{*}=1/\bar{z}_{t} is the reflected flow and its vector field vv^{*} is given by the equation z˙=z˙/z2¯=v(z)¯(z)2.\dot{z}^{*}=-\overline{\dot{z}/z^{2}}=-\overline{v(z)}(z^{*})^{2}. Thus we have v(z)=z2v(1/z¯)¯.v^{*}(z)=-z^{2}\overline{v({1}/{\bar{z}})}. Using a similar method as in the previous corollary, we obtain the following form of Ward’s equations in the radial case.

Corollary 7.8 (Ward’s equations in 𝔻\mathbb{D}).

In the 𝔻\mathbb{D}-uniformization, for any tensor product 𝒳𝛃\mathcal{X}_{\boldsymbol{\beta}} of fields in the extended OPE family 𝛃,\mathcal{F}_{\boldsymbol{\beta}},

(7.9) 2ζ2𝐄T𝜷(ζ)𝒳𝜷=2ζ2𝐄T𝜷(ζ)𝐄𝒳𝜷+𝐄(vζ++vζ)𝒳𝜷.2\zeta^{2}\mathbf{E}\,T_{\boldsymbol{\beta}}(\zeta)\mathcal{X}_{\boldsymbol{\beta}}=2\zeta^{2}\mathbf{E}\,T_{\boldsymbol{\beta}}(\zeta)\,\mathbf{E}\,\mathcal{X}_{\boldsymbol{\beta}}+\mathbf{E}\,(\mathcal{L}_{v_{\zeta}}^{+}+\mathcal{L}_{v_{\zeta^{*}}}^{-})\mathcal{X}_{\boldsymbol{\beta}}.

Let us emphasize that the above formulas apply to the strings without nodes at ζ.\zeta. The case where ζ\zeta is a node is discussed in the following subsection. We use OPE calculus and local operators. Theorem 1.3 now follows from Corollaries 7.7 – 7.8 and the next lemma.

Lemma 7.9.

We have

(7.10) 𝐄T𝜷(z)=𝒫𝜷1kz+𝒫𝜷+𝒫𝜷1kz¯𝒫𝜷\mathbf{E}\,T_{\boldsymbol{\beta}}(z)=\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{+}_{k_{z}}\mathcal{P}_{\boldsymbol{\beta}}+\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{-}_{k_{\bar{z}}}\mathcal{P}_{\boldsymbol{\beta}}

in the identity chart of \mathbb{H} and

(7.11) 2z2𝐄T𝜷(z)=𝒫𝜷1vz+𝒫𝜷+𝒫𝜷1vz𝒫𝜷2z^{2}\mathbf{E}\,T_{\boldsymbol{\beta}}(z)=\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{+}_{v_{z}}\mathcal{P}_{\boldsymbol{\beta}}+\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{-}_{v_{z^{*}}}\mathcal{P}_{\boldsymbol{\beta}}

in the identity chart of 𝔻.\mathbb{D}.

Proof.

Let λk±=λ(βk±),(λ(σ)=σ2/2σb).\lambda_{k}^{\pm}=\lambda(\beta_{k}^{\pm}),(\lambda(\sigma)=\sigma^{2}/2-\sigma b). Since 𝒫𝜷=C(b)[𝜷]\mathcal{P}_{\boldsymbol{\beta}}=C_{(b)}[\boldsymbol{\beta}] is a [λj+,λj][\lambda_{j}^{+},\lambda_{j}^{-}]-differential at qj,q_{j},

𝒫𝜷1kz+𝒫𝜷=jλj+(zqj)2+jlog𝒫𝜷zqj,𝒫𝜷1kz¯𝒫𝜷=jλj(zq¯j)2+¯jlog𝒫𝜷zq¯j\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{+}_{k_{z}}\mathcal{P}_{\boldsymbol{\beta}}=\sum_{j}\frac{\lambda_{j}^{+}}{(z-q_{j})^{2}}+\frac{\partial_{j}\log\mathcal{P}_{\boldsymbol{\beta}}}{z-q_{j}},\qquad\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{-}_{k_{\bar{z}}}\mathcal{P}_{\boldsymbol{\beta}}=\sum_{j}\frac{\lambda_{j}^{-}}{(z-\bar{q}_{j})^{2}}+\frac{{\bar{\partial}}_{j}\log\mathcal{P}_{\boldsymbol{\beta}}}{z-\bar{q}_{j}}

in the identity chart of .\mathbb{H}. Differentiating 𝒫𝜷=C(b)[𝜷],\mathcal{P}_{\boldsymbol{\beta}}=C_{(b)}[\boldsymbol{\beta}], we have

jlog𝒫𝜷=kjβj+βk+qjqk+kβj+βkqjq¯k,¯jlog𝒫𝜷=kjβjβkq¯jq¯k+kβjβk+q¯jqk.\partial_{j}\log\mathcal{P}_{\boldsymbol{\beta}}=\sum_{k\neq j}\frac{\beta_{j}^{+}\beta_{k}^{+}}{q_{j}-q_{k}}+\sum_{k}\frac{\beta_{j}^{+}\beta_{k}^{-}}{q_{j}-\bar{q}_{k}},\qquad{\bar{\partial}}_{j}\log\mathcal{P}_{\boldsymbol{\beta}}=\sum_{k\neq j}\frac{\beta_{j}^{-}\beta_{k}^{-}}{\bar{q}_{j}-\bar{q}_{k}}+\sum_{k}\frac{\beta_{j}^{-}\beta_{k}^{+}}{\bar{q}_{j}-q_{k}}.

On the other hand, it follows from (5.16) and 𝐄T𝜷=12j𝜷2+ibj𝜷\mathbf{E}\,T_{\boldsymbol{\beta}}=-\frac{1}{2}j_{\boldsymbol{\beta}}^{2}+ibj^{\prime}_{\boldsymbol{\beta}} that

𝐄T𝜷(z)\displaystyle\mathbf{E}\,T_{\boldsymbol{\beta}}(z) =j(λj+(zqj)2+λj(zq¯j)2)\displaystyle=\sum_{j}\bigg{(}\frac{\lambda_{j}^{+}}{(z-q_{j})^{2}}+\frac{\lambda_{j}^{-}}{(z-\bar{q}_{j})^{2}}\bigg{)}
+j<k(βj+βk+(zqj)(zqk)+βjβj(zq¯j)(zq¯k))+j,kβj+βk(zqj)(zq¯k).\displaystyle+\sum_{j<k}\bigg{(}\frac{\beta_{j}^{+}\beta_{k}^{+}}{(z-q_{j})(z-q_{k})}+\frac{\beta_{j}^{-}\beta_{j}^{-}}{(z-\bar{q}_{j})(z-\bar{q}_{k})}\bigg{)}+\sum_{j,k}\frac{\beta_{j}^{+}\beta_{k}^{-}}{(z-q_{j})(z-\bar{q}_{k})}.

As a rational function in z,z,

j1zqj\displaystyle\sum_{j}\frac{1}{z-q_{j}} (kjβj+βk+qjqk+kβj+βkqjq¯k)+j1zq¯j(kjβjβkq¯jq¯k+kβjβk+q¯jqk)\displaystyle\bigg{(}\sum_{k\neq j}\frac{\beta_{j}^{+}\beta_{k}^{+}}{q_{j}-q_{k}}+\sum_{k}\frac{\beta_{j}^{+}\beta_{k}^{-}}{q_{j}-\bar{q}_{k}}\bigg{)}+\sum_{j}\frac{1}{z-\bar{q}_{j}}\bigg{(}\sum_{k\neq j}\frac{\beta_{j}^{-}\beta_{k}^{-}}{\bar{q}_{j}-\bar{q}_{k}}+\sum_{k}\frac{\beta_{j}^{-}\beta_{k}^{+}}{\bar{q}_{j}-q_{k}}\bigg{)}
=j<k(βj+βk+(zqj)(zqk)+βjβj(zq¯j)(zq¯k))+j,kβj+βk(zqj)(zq¯k),\displaystyle=\sum_{j<k}\bigg{(}\frac{\beta_{j}^{+}\beta_{k}^{+}}{(z-q_{j})(z-q_{k})}+\frac{\beta_{j}^{-}\beta_{j}^{-}}{(z-\bar{q}_{j})(z-\bar{q}_{k})}\bigg{)}+\sum_{j,k}\frac{\beta_{j}^{+}\beta_{k}^{-}}{(z-q_{j})(z-\bar{q}_{k})},

which shows (7.10). The formula (7.11) can be shown similarly. ∎

7.4. BPZ equations

Recall the definition of Virasoro generators LnLn𝜷L_{n}\equiv L_{n}^{\boldsymbol{\beta}}: they are operators XLnXX\mapsto L_{n}X acting on fields, Ln𝜷X=T𝜷(n2)X,L_{n}^{\boldsymbol{\beta}}X=T_{\boldsymbol{\beta}}*_{(-n-2)}X, i.e.,

(7.12) T𝜷(ζ)X(z)=+(L0X)(z)(ζz)2+(L1X)(z)ζz+(L2X)(z)+(ζz).T_{\boldsymbol{\beta}}(\zeta)X(z)=\cdots+\frac{(L_{0}X)(z)}{(\zeta-z)^{2}}+\frac{(L_{-1}X)(z)}{\zeta-z}+(L_{-2}X)(z)+\cdots\qquad(\zeta\to z).

We write L(z)=L2(z)L(z)=L_{-2}(z) in (,)(\mathbb{H},\infty) and

(7.13) L(z)=2z2L2(z)+3zL1(z)+L0(z)L(z)=2z^{2}L_{-2}(z)+3zL_{-1}(z)+L_{0}(z)

in (𝔻,0).(\mathbb{D},0).

Recall the Belavin-Polyakov-Zamolodchikov equations (BPZ equations [7]) in the chordal case with 𝜷=2bq,\boldsymbol{\beta}=2b\cdot q, see [18, Proposition 5.13].

Proposition 7.10.

Let Y,X1,,Xn𝛃Y,X_{1},\cdots,X_{n}\in\mathcal{F}_{\boldsymbol{\beta}} and let XX be the tensor product of XjX_{j}’s. Then

𝐄(T𝜷Y)(z)X=𝐄Y(z)kz+X+𝐄kz¯[Y(z)X],𝜷=2bq,\mathbf{E}\,(T_{\boldsymbol{\beta}}*Y)(z)\,X=\mathbf{E}\,Y(z)\mathcal{L}^{+}_{k_{z}}X+\mathbf{E}\,\mathcal{L}^{-}_{k_{\bar{z}}}[Y(z)X],\qquad\boldsymbol{\beta}=2b\cdot q,

where all fields are evaluated in the identity chart of (,).(\mathbb{H},\infty).

Somewhat symbolically, we have

L(z)=ˇkz++kz¯,L(z)=\check{\mathcal{L}}^{+}_{k_{z}}+\mathcal{L}^{-}_{k_{\bar{z}}},

where ˇkz+\check{\mathcal{L}}^{+}_{k_{z}} means kz+({z}),\mathcal{L}^{+}_{k_{z}}(\mathbb{H}\setminus\{z\}), e.g., ˇkz+Y(z)X:=Y(z)kz+X.\check{\mathcal{L}}^{+}_{k_{z}}Y(z)X:=Y(z)\mathcal{L}^{+}_{k_{z}}X. We can neglect the check mark if we come to terms that we never differentiate at the poles of v.v.

The previous proposition can be extended to the general background charge, see the next two theorems. We only present the proof of BPZ equations in the radial case. The chordal case can be proved in a similar way.

Theorem 7.11 (BPZ equations in \mathbb{H}).

Let Y,X1,,Xn𝛃Y,X_{1},\cdots,X_{n}\in\mathcal{F}_{\boldsymbol{\beta}} and let XX be the tensor product of XjX_{j}’s. Then

𝐄(T𝜷Y)(z)X=𝐄Y(z)𝒫𝜷1kz+𝒫𝜷X+𝐄𝒫𝜷1kz¯[𝒫𝜷Y(z)X],\mathbf{E}\,(T_{\boldsymbol{\beta}}*Y)(z)\,X=\mathbf{E}\,Y(z)\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{+}_{k_{z}}\mathcal{P}_{\boldsymbol{\beta}}X+\mathbf{E}\,\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{-}_{k_{\bar{z}}}[\mathcal{P}_{\boldsymbol{\beta}}Y(z)X],

where all fields are evaluated in the identity chart of .\mathbb{H}.

Theorem 7.12 (BPZ equations in 𝔻\mathbb{D}).

Let Y,X1,,Xn𝛃Y,X_{1},\cdots,X_{n}\in\mathcal{F}_{\boldsymbol{\beta}} and let XX be the tensor product of XjX_{j}’s. Then

𝐄L(z)Y(z)X=𝐄Y(z)𝒫𝜷1vz+𝒫𝜷X+𝐄𝒫𝜷1vz[𝒫𝜷Y(z)X],\mathbf{E}\,L(z)Y(z)\,X=\mathbf{E}\,Y(z)\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{+}_{v_{z}}\mathcal{P}_{\boldsymbol{\beta}}X+\mathbf{E}\,\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{-}_{v_{z^{*}}}[\mathcal{P}_{\boldsymbol{\beta}}Y(z)X],

where all fields are evaluated in the identity chart of 𝔻\mathbb{D} and L(z)L(z) is given by (7.13).

Proof.

Recall Ward’s OPE, SingζzT𝜷(ζ)Y(z)=kζ+Y(z)\operatorname{Sing}_{\zeta\to z}T_{\boldsymbol{\beta}}(\zeta)Y(z)=\mathcal{L}^{+}_{k_{\zeta}}Y(z) (see [18, Proposition 5.3]). It follows immediately that

(7.14) 2ζ2SingζzT𝜷(ζ)Y(z)=2ζ2kζ+Y(z).2\zeta^{2}\operatorname{Sing}_{\zeta\to z}T_{\boldsymbol{\beta}}(\zeta)Y(z)=\mathcal{L}^{+}_{2\zeta^{2}k_{\zeta}}Y(z).

Denote δζ(η)=2ζ2kζ(η)vζ(η)=2ζ+η.\delta_{\zeta}(\eta)=2\zeta^{2}k_{\zeta}(\eta)-v_{\zeta}(\eta)=2\zeta+\eta. By (7.12), we observe that

δζ+Y(z)=12πi(z)δζ(η)T𝜷(η)Y(z)dη=L0Y(z)+(2ζ+z)L1Y(z).\mathcal{L}_{\delta_{\zeta}}^{+}Y(z)=\frac{1}{2\pi i}\oint_{(z)}\delta_{\zeta}(\eta)T_{\boldsymbol{\beta}}(\eta)Y(z)\,\mathrm{d}\eta=L_{0}Y(z)+(2\zeta+z)L_{-1}Y(z).

It follows from (7.14) that

(7.15) 2ζ2SingζzT𝜷(ζ)Y(z)vζ+Y(z)=L0Y(z)+(2ζ+z)L1Y(z).2\zeta^{2}\operatorname{Sing}_{\zeta\to z}T_{\boldsymbol{\beta}}(\zeta)Y(z)-\mathcal{L}^{+}_{v_{\zeta}}Y(z)=L_{0}Y(z)+(2\zeta+z)L_{-1}Y(z).

On the other hand, by Theorem 1.3,

2ζ2𝐄T𝜷(ζ)Y(z)X=𝐄𝒫𝜷1vζ+𝒫𝜷Y(z)X+𝐄𝒫𝜷1vζ𝒫𝜷Y(z)X,2\zeta^{2}\,\mathbf{E}\,T_{\boldsymbol{\beta}}(\zeta)Y(z)X=\mathbf{E}\,\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{+}_{v_{\zeta}}\mathcal{P}_{\boldsymbol{\beta}}Y(z)X+\mathbf{E}\,\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{-}_{v_{\zeta^{*}}}\mathcal{P}_{\boldsymbol{\beta}}Y(z)X,

where the first term on the right-hand side is

𝐄𝒫𝜷1vζ+𝒫𝜷Y(z)X=𝐄vζ+Y(z)X+𝐄Y(z)𝒫𝜷1vζ+𝒫𝜷X.\mathbf{E}\,\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{+}_{v_{\zeta}}\mathcal{P}_{\boldsymbol{\beta}}Y(z)X=\mathbf{E}\,\mathcal{L}^{+}_{v_{\zeta}}Y(z)X+\mathbf{E}\,Y(z)\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{+}_{v_{\zeta}}\mathcal{P}_{\boldsymbol{\beta}}X.

Combining the last three equations and subtracting all singular terms, we get

𝐄(2ζ2Regζz(T𝜷(ζ)Y(z))X\displaystyle\mathbf{E}\,\big{(}2\zeta^{2}\operatorname{Reg}_{\zeta\to z}(T_{\boldsymbol{\beta}}(\zeta)Y(z))X +L0Y(z)X+(2ζ+z)L1Y(z)X)\displaystyle+L_{0}Y(z)X+(2\zeta+z)L_{-1}Y(z)X\big{)}
=𝐄Y(z)𝒫𝜷1vζ+𝒫𝜷X+𝐄𝒫𝜷1vζ𝒫𝜷Y(z)X,\displaystyle=\mathbf{E}\,Y(z)\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{+}_{v_{\zeta}}\mathcal{P}_{\boldsymbol{\beta}}X+\mathbf{E}\,\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{-}_{v_{\zeta^{*}}}\mathcal{P}_{\boldsymbol{\beta}}Y(z)X,

where RegζzT𝜷(ζ)Y(z)\operatorname{Reg}_{\zeta\to z}T_{\boldsymbol{\beta}}(\zeta)Y(z) is the regular part of the operator product expansion. Theorem now follows by taking the limit ζz\zeta\to z in both sides. For instance, the left-hand side converges to 𝐄L(z)Y(z)X.\mathbf{E}\,L(z)Y(z)\,X.

7.5. Null vectors

After we briefly review the level two degeneracy equations for current primary fields, we apply these equations and the BPZ equations to the (tensor/OPE) products of the one-leg operators and the fields in the extended OPE family.

We denote by {Jn}\{J_{n}\} and {Ln}\{L_{n}\} the current generators and the Virasoro generators, the modes of J𝜷J_{\boldsymbol{\beta}} and T𝜷T_{\boldsymbol{\beta}} in 𝜷\mathcal{F}_{\boldsymbol{\beta}} theory, respectively:

Jn(z):=12πi(z)(ζz)nJ𝜷(ζ)dζ,Ln(z):=12πi(z)(ζz)n+1T𝜷(ζ)dζ.J_{n}(z):=\frac{1}{2\pi i}\oint_{(z)}(\zeta-z)^{n}J_{\boldsymbol{\beta}}(\zeta)\,\mathrm{d}\zeta,\qquad L_{n}(z):=\frac{1}{2\pi i}\oint_{(z)}(\zeta-z)^{n+1}T_{\boldsymbol{\beta}}(\zeta)\,\mathrm{d}\zeta.

As operators acting on fields in 𝜷,\mathcal{F}_{\boldsymbol{\beta}}, we have the following equations:

[Jm,Jn]=nδm+n,0,[Lm,Jn]=nJm+n+ibm(m+1)δm+n,0,[J_{m},J_{n}]=n\delta_{m+n,0},\quad[L_{m},J_{n}]=-nJ_{m+n}+ibm(m+1)\delta_{m+n,0},

and

[Lm,Ln]=(mn)Lm+n+c12m(m21)δm+n,0,[L_{m},L_{n}]=(m-n)L_{m+n}+\frac{c}{12}m(m^{2}-1)\delta_{m+n,0},

where cc is the central charge, c=112b2.c=1-12b^{2}.

By definition, X𝜷X\in\mathcal{F}_{\boldsymbol{\beta}} is (Virasoro) primary if XX is a [λ,λ][\lambda,\lambda_{*}]-differential; equivalently

(7.16) L1X=0,L0X=λX,L1X=X,L_{\geq 1}X=0,\qquad L_{0}X=\lambda X,\qquad L_{-1}X=\partial X,

and similar equations hold for X¯.\bar{X}. Here LkX=0L_{\geq k}X=0 means that LnX=0L_{n}X=0 for all nk.n\geq k. A (Virasoro) primary field XX is called current primary with charges σ\sigma and σ\sigma_{*} if J1X=J1X¯=0,J_{\geq 1}X=J_{\geq 1}\bar{X}=0, and

J0X=iσX,J0X¯=iσ¯X¯.J_{0}X=-i\sigma X,\quad J_{0}\bar{X}=i\bar{\sigma}_{*}\bar{X}.

Let us recall the characterization of level two degenerate current primary fields.

Proposition 7.13 (Proposition 11.2 in [18]).

Let 𝒪\mathcal{O} be a current primary field in 𝛃,\mathcal{F}_{\boldsymbol{\beta}}, and let σ,σ\sigma,\sigma_{*} be charges of 𝒪.\mathcal{O}. If 2σ(b+σ)=1,η=1/(2σ2),2\sigma(b+\sigma)=1,\,\eta=-1/(2\sigma^{2}), then

(L2+ηL12)𝒪=0.(L_{-2}+\eta L_{-1}^{2})\mathcal{O}=0.

If the parameters aa and bb are related to the SLE parameter κ\kappa as

a=±2/κ,b=a(κ/41),a=\pm\sqrt{2/\kappa},\qquad b=a(\kappa/4-1),

then 2a(a+b)=12a(a+b)=1 and the formal fields 𝒪(a),𝒪(ab)\mathcal{O}^{(a)},\mathcal{O}^{(-a-b)} satisfy

(7.17) L2𝒪(a)=κ4z2𝒪(a),L2𝒪(ab)=4κz2𝒪(ab)L_{-2}\,\mathcal{O}^{(a)}=\frac{\kappa}{4}\,\partial_{z}^{2}\mathcal{O}^{(a)},\qquad L_{-2}\,\mathcal{O}^{(-a-b)}=\frac{4}{\kappa}\,\partial_{z}^{2}\mathcal{O}^{(-a-b)}

(in all charts). Since 𝒪(a),𝒪(ab)\mathcal{O}^{(a)},\mathcal{O}^{(-a-b)} are formal fields, we need to assume the neutrality condition (NC0)(\mathrm{NC}_{0}) on 𝒪(a)X,𝒪(ab)X\mathcal{O}^{(a)}X,\mathcal{O}^{(-a-b)}X for some product XX of formal fields. For example, the one-leg operator Λp\Lambda_{p} (both in the chordal case and in the radial one) satisfy the level two degeneracy equation. As mentioned before, we will use this level two degeneracy equation for Λp\Lambda_{p} to establish the connection between the chordal/radial SLE theory and conformal field theory.

• Standard chordal case with 𝛃=2bq\boldsymbol{\beta}=2b\cdot q.  Recall the BPZ equations in (,):(\mathbb{H},\infty):

L(z)=ˇkz++kz¯,kζ(η)=1ζη,L(z)=\check{\mathcal{L}}^{+}_{k_{z}}+\mathcal{L}^{-}_{k_{\bar{z}}},\qquad k_{\zeta}(\eta)=\frac{1}{\zeta-\eta},

which we apply to the field Y(z)XY(z)X in 𝜷,\mathcal{F}_{\boldsymbol{\beta}}, where XX has no node at z.z. In all applications YY is holomorphic, so we have

(7.18) (L(z)Y(z))X=Y(z)(ˇkz++kz¯)X=Y(z)(kz++kz¯)X.\big{(}L(z)Y(z)\big{)}X=Y(z)(\check{\mathcal{L}}^{+}_{k_{z}}+\mathcal{L}^{-}_{k_{\bar{z}}})X=Y(z)(\mathcal{L}^{+}_{k_{z}}+\mathcal{L}^{-}_{k_{\bar{z}}})X.

If XX is also holomorphic (in particular, in the “boundary” situation), then (7.18) becomes

(L(z)Y(z))X=Y(z)kz+X.\big{(}L(z)Y(z)\big{)}X=Y(z)\mathcal{L}^{+}_{k_{z}}X.

Let us discuss both sides in the equation (7.18).

We study the cases that YY is the null vector

Y(z)=𝒪(a)(z),Y(z)=𝒪(ab)(z).Y(z)=\mathcal{O}^{(a)}(z),\qquad Y(z)=\mathcal{O}^{(-a-b)}(z).

Of course, YY is a formal field, so we need to assume the neutrality condition (NC0)(\mathrm{NC}_{0}) in Y(z)X.Y(z)X. Recall that the conformal dimensions h,hh,h^{\prime} of 𝒪(a),𝒪(ab),\mathcal{O}^{(a)},\mathcal{O}^{(-a-b)}, respectively at zz are

(7.19) h:=3a2212=6κ2κ,h:=38a212=3κ816.h:=\frac{3a^{2}}{2}-\frac{1}{2}=\frac{6-\kappa}{2\kappa},\qquad h^{\prime}:=\frac{3}{8a^{2}}-\frac{1}{2}=\frac{3\kappa-8}{16}.

Recall the level two degeneracy equation (7.17) for the formal field 𝒪(a)\mathcal{O}^{(a)}. In the boundary case z=xz=x\in\mathbb{R} we have

(7.20) L2𝒪(a)(x)=κ4x2𝒪(a)(x)L_{-2}\,\mathcal{O}^{(a)}(x)=\frac{\kappa}{4}\,\partial_{x}^{2}\mathcal{O}^{(a)}(x)

in the (,)(\mathbb{H},\infty)-uniformization because x=+¯\partial_{x}=\partial+{\bar{\partial}} and 𝒪(a)\mathcal{O}^{(a)} is holomorphic.

In the case that X=𝒪𝜷[𝝉]X=\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}] with 𝝉=τjzj,\boldsymbol{\tau}=\sum\tau_{j}\cdot z_{j}, we have

Y(z)kz+X=(kz(zj)zj+λjzjkz(zj))Y(z)X,Y(z)\mathcal{L}^{+}_{k_{z}}X=\big{(}\sum k_{z}(z_{j})\partial_{z_{j}}+\lambda_{j}\partial_{z_{j}}k_{z}(z_{j})\big{)}Y(z)X,

where λj=λb(τj)=12τj2bτj.\lambda_{j}=\lambda_{b}(\tau_{j})=\frac{1}{2}\tau_{j}^{2}-b\tau_{j}.

• Standard radial case with 𝛃=bq+bq\boldsymbol{\beta}=b\cdot q+b\cdot q^{*}.  We apply the BPZ equation to the tensor product Y(z)XY(z)X of fields in 𝜷(D,q)\mathcal{F}_{\boldsymbol{\beta}}(D,q) with holomorphic Y:Y:

(7.21) (L(z)Y(z))X=Y(z)𝒫q1(vz++vz)𝒫qX,vζ(η)=ηζ+ηζη.\big{(}L(z)Y(z)\big{)}X=Y(z)\mathcal{P}_{q}^{-1}(\mathcal{L}^{+}_{v_{z}}+\mathcal{L}^{-}_{v_{z^{*}}})\mathcal{P}_{q}X,\qquad v_{\zeta}(\eta)=\eta\frac{\zeta+\eta}{\zeta-\eta}.

Let us discuss both sides in the equation (7.21).

We consider the (𝔻,0)(\mathbb{D},0)-uniformization. For Y=𝒪(a),Y=\mathcal{O}^{(a)}, we have

(7.22) L(z)𝒪(a)=(κ2z2z2+3zz+h)𝒪(a),L(z)\mathcal{O}^{(a)}=\Big{(}\frac{\kappa}{2}\,z^{2}\partial_{z}^{2}+3z\partial_{z}+h\Big{)}\mathcal{O}^{(a)},

where hh is given by (7.19). In particular, for z=eiθ𝔻,z=\mathrm{e}^{i\theta}\in\partial\mathbb{D},

L(eiθ)𝒪(a)=(κ(12θ2+ihθ)+h)𝒪(a).L(\mathrm{e}^{i\theta})\mathcal{O}^{(a)}=\Big{(}\!-\kappa\big{(}\frac{1}{2}\partial_{\theta}^{2}+ih\partial_{\theta}\big{)}+h\Big{)}\mathcal{O}^{(a)}.

In the case of O(ab)O^{(-a-b)} the corresponding differential operator is

L(eiθ)=κ(12θ2+ihθ)+h,L(\mathrm{e}^{i\theta})=-\kappa^{\prime}\big{(}\frac{1}{2}\partial_{\theta}^{2}+ih^{\prime}\partial_{\theta}\big{)}+h^{\prime},

where κ=16/κ\kappa^{\prime}=16/\kappa and hh^{\prime} is given by (7.19).

By Lemma 7.9, the right hand-side becomes

Y(z)(vz++vz)X+2z2Y(z)X𝐄T𝜷(z).Y(z)(\mathcal{L}^{+}_{v_{z}}+\mathcal{L}^{-}_{v_{z^{*}}})X+2z^{2}Y(z)X\,\mathbf{E}\,T_{\boldsymbol{\beta}}(z).

8. Connection to the theory of forward chordal/radial SLE with forces and spins

In Subsections 8.1 – 8.2 we define the one-leg operators for chordal/radial SLEs with forces and spins as the OPE exponentials with specific background charge 𝜷\boldsymbol{\beta} at the marked points. The insertion of a one-leg operator changes the boundary values of Fock space fields. As the PDEs for correlation functions of the fields in the extended OPE family 𝜷(D),\mathcal{F}_{\boldsymbol{\beta}}(D), the BPZ-Cardy equations for such insertion operations are used to prove the connection (Theorem 1.5) between the theory 𝜷\mathcal{F}_{\boldsymbol{\beta}} and the chordal/radial SLE[𝜷]\mathrm{SLE}[\boldsymbol{\beta}] theory. A dipolar theory studied in [19] is the special of this theory since it is well known ([33, 37]) that dipolar SLE(κ)\mathrm{SLE}(\kappa) with two force points q±q_{\pm} is equivalent to chordal SLE(κ,𝝆)\mathrm{SLE}(\kappa,\boldsymbol{\rho}) with 𝝆=12(κ6)q++12(κ6)q.\boldsymbol{\rho}=\frac{1}{2}(\kappa-6)\cdot q_{+}+\frac{1}{2}(\kappa-6)\cdot q_{-}. In Subsection 8.6 we present field theoretic approach to chordal SLE(κ,𝝆)\mathrm{SLE}(\kappa,\boldsymbol{\rho}) restriction martingale-observable introduced in [13, Lemma 4].

8.1. One-leg operators for chordal SLEs with forces and spins

Let (D,p,q)(D,p,q) be a simply-connected domain with two distinct marked boundary points p,qD,p,q\in\partial D, and 𝜷\boldsymbol{\beta} be a symmetric background charge on S=DdoubleS=D^{\mathrm{double}} with

𝜷=ap+𝜼,psupp𝜼\boldsymbol{\beta}=a\cdot p+\boldsymbol{\eta},\quad p\notin\mathrm{supp}\,\boldsymbol{\eta}

satisfying the neutrality condition (NCb):(\mathrm{NC}_{b}):

𝜷=2b.\int\boldsymbol{\beta}=2b.

Here the parameters aa and bb are related to the SLE parameter κ\kappa as

a=±2/κ,b=a(κ/41).a=\pm\sqrt{2/\kappa},\qquad b=a(\kappa/4-1).

Let 𝝉=apaq.\boldsymbol{\tau}=a\cdot p-a\cdot q. For any divisor 𝝈,\boldsymbol{\sigma}, we denote 𝝈ˇ=𝝈𝝉\check{\boldsymbol{\sigma}\,}=\boldsymbol{\sigma}-\boldsymbol{\tau} and 𝝈^=𝝈+𝝉.\hat{\boldsymbol{\sigma}\,}=\boldsymbol{\sigma}+\boldsymbol{\tau}.

• Insertions.  The insertion of the one-leg operator

ΛpΛ(p):=𝒪𝜷ˇ[𝝉]\Lambda_{p}\equiv\Lambda(p):=\mathcal{O}_{\check{\boldsymbol{\beta}\,}}[\boldsymbol{\tau}]

produces an operator 𝒳𝒳^\mathcal{X}\mapsto\widehat{\mathcal{X}} acting on Fock space functionals/fields by the rules (5.20) and the formula

Φ^𝜷ˇ(z)=Φ𝜷ˇ(z)+2aargw(z)=Φ𝜷(z),\widehat{\Phi}_{\check{\boldsymbol{\beta}\,}}(z)=\Phi_{\check{\boldsymbol{\beta}\,}}(z)+2a\arg w(z)=\Phi_{\boldsymbol{\beta}}(z),

where w:(D,p,q)(,0,)w:(D,p,q)\to(\mathbb{H},0,\infty) is a conformal map.

We denote

𝐄^𝒳:=𝐄Λp𝒳𝐄Λp=𝐄Λpeff𝒳𝐄Λpeff=𝐄𝒳eiaΦ+(p,q),\widehat{\mathbf{E}}\,\mathcal{X}:=\frac{\mathbf{E}\,\Lambda_{p}\mathcal{X}}{\mathbf{E}\,\Lambda_{p}}=\frac{\mathbf{E}\,\Lambda_{p}^{\mathrm{eff}}\mathcal{X}}{\mathbf{E}\,\Lambda_{p}^{\mathrm{eff}}}=\mathbf{E}\,\mathcal{X}\mathrm{e}^{\odot ia\Phi^{+}(p,q)},

where Λpeff\Lambda_{p}^{\mathrm{eff}} is the effective one-leg operator

Λpeff=𝒫𝜷ˇΛp=C(b)[𝜷ˇ]𝒪𝜷ˇ[𝝉]=C(b)[𝜷]V[𝝉]=V𝜷ˇ[𝜷].\Lambda_{p}^{\mathrm{eff}}=\mathcal{P}_{\check{\boldsymbol{\beta}\,}}\Lambda_{p}=C_{(b)}[\check{\boldsymbol{\beta}\,}]\mathcal{O}_{\check{\boldsymbol{\beta}\,}}[\boldsymbol{\tau}]=C_{(b)}[\boldsymbol{\beta}]V^{\odot}[\boldsymbol{\tau}]=V_{\check{\boldsymbol{\beta}\,}}[\boldsymbol{\beta}].

Theorem 1.1 says that

𝐄^𝒳=𝐄𝒳^\widehat{\mathbf{E}}\,\mathcal{X}=\mathbf{E}\,\widehat{\mathcal{X}}

for any string 𝒳\mathcal{X} of fields in 𝜷ˇ.\mathcal{F}_{\check{\boldsymbol{\beta}\,}}. We define the partition function associated with a symmetric background charge 𝜷\boldsymbol{\beta} by

Z𝜷:=|C(b)[𝜷]|.Z_{\boldsymbol{\beta}}:=\big{|}C_{(b)}[\boldsymbol{\beta}]\big{|}.

In the \mathbb{H}-uniformization, C(b)[𝜷]C_{(b)}[\boldsymbol{\beta}] is non-negative (up to a phase), see Example (b) in Subsection 2.5 and thus we have Z𝜷=C(b)[𝜷].Z_{\boldsymbol{\beta}}=C_{(b)}[\boldsymbol{\beta}].

• BPZ equations.  Let ξD\xi\in\partial D and 𝜷ξ=𝜷ap+aξ.\boldsymbol{\beta}_{\xi}=\boldsymbol{\beta}-a\cdot p+a\cdot\xi. Let XX be the tensor product X=X1(z1)Xn(zn)X=X_{1}(z_{1})\cdots X_{n}(z_{n}) of fields XjX_{j} in 𝜷ˇ\mathcal{F}_{\check{\boldsymbol{\beta}\,}} (zjD).(z_{j}\in D). It follows from Theorem 7.11 that

(8.1) 12a2ξ2𝐄ΛξeffX=𝐄ˇkξΛξeffX,kξ(z):=1ξz,\frac{1}{2a^{2}}\partial_{\xi}^{2}\mathbf{E}\Lambda_{\xi}^{\mathrm{eff}}X=\mathbf{E}\,\check{\mathcal{L}}_{k_{\xi}}\Lambda_{\xi}^{\mathrm{eff}}X,\qquad k_{\xi}(z):=\frac{1}{\xi-z},

in the \mathbb{H}-uniformization. Here ξ=+¯\partial_{\xi}=\partial+\bar{\partial} is the operator of differentiation with respect to the real variable ξ\xi and ˇkξ\check{\mathcal{L}}_{k_{\xi}} is taken over the finite notes of XX and supp𝜷ξ{ξ}.\mathrm{supp}\,\boldsymbol{\beta}_{\xi}\setminus\{\xi\}. In particular, the partition function ZξZ𝜷ξZ_{\xi}\equiv Z_{\boldsymbol{\beta}_{\xi}} satisfies the null vector equation

(8.2) 12a2ξ2Zξ=ˇkξZξ,\frac{1}{2a^{2}}\partial_{\xi}^{2}Z_{\xi}=\check{\mathcal{L}}_{k_{\xi}}Z_{\xi},

in the \mathbb{H}-uniformization since Zξ=𝐄ΛξeffZ_{\xi}=\mathbf{E}\Lambda_{\xi}^{\mathrm{eff}} up to a phase in the \mathbb{H}-uniformization. Here ˇkξ\check{\mathcal{L}}_{k_{\xi}} is taken over supp𝜷ξ{ξ}.\mathrm{supp}\,\boldsymbol{\beta}_{\xi}\setminus\{\xi\}.

• BPZ-Cardy equations.  For ξ\xi\in\mathbb{R} and a tensor product X=X1(z1)Xn(zn)X=X_{1}(z_{1})\cdots X_{n}(z_{n}) of fields XjX_{j} in 𝜷ˇ\mathcal{F}_{\check{\boldsymbol{\beta}\,}} (zj)(z_{j}\in\mathbb{H}), we denote

Rξ(𝒛,𝒒)𝐄^ξX:=𝐄ΛξX𝐄Λξ=𝐄ΛξeffX𝐄Λξeff=𝐄XeiaΦ+(ξ,q).R_{\xi}(\boldsymbol{z},\boldsymbol{q})\equiv\widehat{\mathbf{E}}_{\xi}X:=\frac{\mathbf{E}\,\Lambda_{\xi}X}{\mathbf{E}\,\Lambda_{\xi}}=\frac{\mathbf{E}\,\Lambda_{\xi}^{\mathrm{eff}}X}{\mathbf{E}\,\Lambda_{\xi}^{\mathrm{eff}}}=\mathbf{E}\,X\mathrm{e}^{\odot ia\Phi^{+}(\xi,q)}.

The following is the first half of Theorem 1.4.

Proposition 8.1 (BPZ-Cardy equations in \mathbb{H}).

In the identity chart of ,\mathbb{H}, we have

(8.3) 12a2(ξ2𝐄^ξX+2(ξlogZξ)ξ𝐄^ξX)=ˇkξ𝐄^ξX,kξ(z):=1ξz,\frac{1}{2a^{2}}\big{(}\partial_{\xi}^{2}\widehat{\mathbf{E}}_{\xi}\,X+2(\partial_{\xi}\log Z_{\xi})\partial_{\xi}\widehat{\mathbf{E}}_{\xi}\,X\big{)}=\check{\mathcal{L}}_{k_{\xi}}\widehat{\mathbf{E}}_{\xi}\,X,\qquad k_{\xi}(z):=\frac{1}{\xi-z},

where ξ=+¯\partial_{\xi}=\partial+{\bar{\partial}} and ˇkξ\check{\mathcal{L}}_{k_{\xi}} is taken over the finite notes of XX and supp𝛃ξ{ξ}.\mathrm{supp}\,\boldsymbol{\beta}_{\xi}\setminus\{\xi\}.

Proof.

By differentiation,

ξ2𝐄ΛξeffX𝐄Λξeff=ξ2𝐄^ξX+2ξZξZξξ𝐄^ξX+ξ2ZξZξ𝐄^ξX.\frac{\partial_{\xi}^{2}\mathbf{E}\,\Lambda_{\xi}^{\mathrm{eff}}X}{\mathbf{E}\,\Lambda_{\xi}^{\mathrm{eff}}}=\partial_{\xi}^{2}\widehat{\mathbf{E}}_{\xi}\,X+2\frac{\partial_{\xi}Z_{\xi}}{Z_{\xi}}\partial_{\xi}\widehat{\mathbf{E}}_{\xi}\,X+\frac{\partial_{\xi}^{2}Z_{\xi}}{Z_{\xi}}\widehat{\mathbf{E}}_{\xi}\,X.

On the other hand, it follows from the BPZ equations (8.1) that the left-hand side of the above becomes 2a2𝐄ˇkξΛξeffX/𝐄Λξeff.2a^{2}{\mathbf{E}\,\check{\mathcal{L}}_{k_{\xi}}\Lambda_{\xi}^{\mathrm{eff}}X}/{\mathbf{E}\,\Lambda_{\xi}^{\mathrm{eff}}}. By Leibniz’s rule for Lie derivatives, we have

𝐄ˇkξΛξeffX𝐄Λξeff=ˇkξ𝐄^ξX+ˇkξZξZξ𝐄^ξX.\frac{\mathbf{E}\,\check{\mathcal{L}}_{k_{\xi}}\Lambda_{\xi}^{\mathrm{eff}}X}{\mathbf{E}\,\Lambda_{\xi}^{\mathrm{eff}}}=\check{\mathcal{L}}_{k_{\xi}}\widehat{\mathbf{E}}_{\xi}\,X+\frac{\check{\mathcal{L}}_{k_{\xi}}Z_{\xi}}{Z_{\xi}}\,\widehat{\mathbf{E}}_{\xi}\,X.

Combining all of the above, we find

12a2(ξ2𝐄^ξX+2ξZξZξξ𝐄^ξX+ξ2ZξZξ𝐄^ξX)=ˇkξ𝐄^ξX+ˇkξZξZξ𝐄^ξX.\frac{1}{2a^{2}}\big{(}\partial_{\xi}^{2}\widehat{\mathbf{E}}_{\xi}\,X+2\frac{\partial_{\xi}Z_{\xi}}{Z_{\xi}}\partial_{\xi}\widehat{\mathbf{E}}_{\xi}\,X+\frac{\partial_{\xi}^{2}Z_{\xi}}{Z_{\xi}}\widehat{\mathbf{E}}_{\xi}\,X\big{)}=\check{\mathcal{L}}_{k_{\xi}}\widehat{\mathbf{E}}_{\xi}\,X+\frac{\check{\mathcal{L}}_{k_{\xi}}Z_{\xi}}{Z_{\xi}}\,\widehat{\mathbf{E}}_{\xi}\,X.

We remark that the coefficient of 𝐄^ξX\widehat{\mathbf{E}}_{\xi}\,X vanishes since the partition function ZξZ_{\xi} satisfies the null vector equation (8.2). ∎

8.2. One-leg operators for radial SLEs with forces and spins

Let (D,p,q)(D,p,q) be a simply-connected domain DD with a marked boundary point pD,p\in\partial D, and a marked interior point qD.q\in D. As in the previous subsection, we consider a symmetric background charge 𝜷\boldsymbol{\beta} on S=DdoubleS=D^{\mathrm{double}} with

𝜷=ap+𝜼,psupp𝜼\boldsymbol{\beta}=a\cdot p+\boldsymbol{\eta},\quad p\notin\mathrm{supp}\,\boldsymbol{\eta}

satisfying the neutrality condition (NCb).(\mathrm{NC}_{b}).

Let η.\eta\in\mathbb{R}. We introduce the one-leg operator Λps\Lambda_{p}^{s} with spin s=iηa2/2s=i\eta{a^{2}}/2 as

Λps𝒪𝜷ˇ[𝝉],𝝉=apa+iδ2qaiδ2q,δ=ηa,\Lambda_{p}^{s}\equiv\mathcal{O}_{\check{\boldsymbol{\beta}\,}}[\boldsymbol{\tau}],\qquad\boldsymbol{\tau}=a\cdot p-\frac{a+i\delta}{2}\cdot q-\frac{a-i\delta}{2}\cdot q^{*},\qquad\delta=\eta a,

where 𝜷ˇ=𝜷𝝉\check{\boldsymbol{\beta}\,}=\boldsymbol{\beta}-\boldsymbol{\tau} is a symmetric background charge.

• Insertions.  As in the chordal theory, the insertion of the one-leg operator Λps\Lambda_{p}^{s} produces an operator 𝒳𝒳^\mathcal{X}\mapsto\widehat{\mathcal{X}} acting on Fock space functionals/fields by the rules (5.20) and the formula

Φ^𝜷ˇ(z)=Φ𝜷(z).\widehat{\Phi}_{\check{\boldsymbol{\beta}\,}}(z)=\Phi_{\boldsymbol{\beta}}(z).

We denote

𝐄^[𝒳]:=𝐄[Λps𝒳]𝐄[Λps]=𝐄[(Λps)eff𝒳]𝐄[(Λps)eff]=𝐄𝒳eiΦ[𝝉],\widehat{\mathbf{E}}[\mathcal{X}]:=\frac{\mathbf{E}[\Lambda_{p}^{s}\mathcal{X}]}{\mathbf{E}[\Lambda_{p}^{s}]}=\frac{\mathbf{E}[(\Lambda_{p}^{s})^{\mathrm{eff}}\mathcal{X}]}{\mathbf{E}[(\Lambda_{p}^{s})^{\mathrm{eff}}]}=\mathbf{E}\,\mathcal{X}\mathrm{e}^{\odot i\Phi[\boldsymbol{\tau}]},

where (Λps)eff(\Lambda_{p}^{s})^{\mathrm{eff}} is the effective one-leg operator with spin ss: (Λps)eff=𝒫𝜷ˇΛps.(\Lambda_{p}^{s})^{\mathrm{eff}}=\mathcal{P}_{\check{\boldsymbol{\beta}\,}}\Lambda_{p}^{s}. Then we have 𝐄^[𝒳]=𝐄[𝒳^].\widehat{\mathbf{E}}[\mathcal{X}]=\mathbf{E}[\widehat{\mathcal{X}}].

Examples.

We consider the case

𝜷s=ap+(ba+iδ2)q+(baiδ2)q,δ=ηa.\boldsymbol{\beta}^{s}=a\cdot p+\Big{(}b-\frac{a+i\delta}{2}\Big{)}\cdot q+\Big{(}b-\frac{a-i\delta}{2}\Big{)}\cdot q^{*},\qquad\delta=\eta a.

Let 𝜷=𝜷0.\boldsymbol{\beta}=\boldsymbol{\beta}^{0}.

(a) The bosonic Φ𝜷s\Phi_{\boldsymbol{\beta}^{s}} is a pre-pre-Schwarzian form of order (ib,ib),(ib,-ib),

Φ𝜷s=Φ𝜷δlog|w|,\Phi_{\boldsymbol{\beta}^{s}}=\Phi_{\boldsymbol{\beta}}-\delta\log|w|,

where w:(D,p,q)(𝔻,1,0)w:(D,p,q)\to(\mathbb{D},1,0) is a conformal map;

(b) The current J𝜷sJ_{\boldsymbol{\beta}^{s}} is a pre-Schwarzian form of order ib,ib,

J𝜷s=J𝜷δ2ww;J_{\boldsymbol{\beta}^{s}}=J_{\boldsymbol{\beta}}-\frac{\delta}{2}\frac{w^{\prime}}{w};

(c) The Virasoro field T𝜷sT_{\boldsymbol{\beta}^{s}} is a Schwarzian form of order 112c,\frac{1}{12}c,

T𝜷s\displaystyle T_{\boldsymbol{\beta}^{s}} =12J𝜷sJ𝜷s+ibJ𝜷s\displaystyle=-\dfrac{1}{2}J_{\boldsymbol{\beta}^{s}}*J_{\boldsymbol{\beta}^{s}}+ib\partial J_{\boldsymbol{\beta}^{s}}
=Aj𝜷sJ+ibJ+c12Sw+h1,2w2w(1w)2+h0,1/2sw2w2+sw2w(1w),\displaystyle=A-j_{\boldsymbol{\beta}^{s}}J+ib\partial J+\frac{c}{12}S_{w}+h_{1,2}\frac{w^{\prime 2}}{w(1-w)^{2}}+h_{0,1/2}^{s}\frac{w^{\prime 2}}{w^{2}}+s\frac{w^{\prime 2}}{w(1-w)},

where j𝜷s=𝐄J𝜷s,j_{\boldsymbol{\beta}^{s}}=\mathbf{E}\,J_{\boldsymbol{\beta}^{s}}, h1,2=12a2abh_{1,2}=\frac{1}{2}a^{2}-ab and h0,1/2s=18(a+iδ)212b2.h_{0,1/2}^{s}=\frac{1}{8}(a+i\delta)^{2}-\frac{1}{2}b^{2}.

Let XX be a tensor product of fields in the extended family 𝜷ˇ(D).\mathcal{F}_{\check{\boldsymbol{\beta}\,}}(D). We assume that ζ\zeta is not a node of X.X. Denote

Rζ𝐄^ζX=𝐄(Λζs)effX𝐄(Λζs)eff,Cζ=𝐄(Λζs)eff=C(b)[𝜷ζ],Zζ=|Cζ|,R_{\zeta}\equiv\widehat{\mathbf{E}}_{\zeta}X=\frac{\mathbf{E}\,(\Lambda^{s}_{\zeta})^{\mathrm{eff}}X}{\mathbf{E}\,(\Lambda^{s}_{\zeta})^{\mathrm{eff}}},\quad C_{\zeta}=\mathbf{E}\,(\Lambda^{s}_{\zeta})^{\mathrm{eff}}=C_{(b)}[\boldsymbol{\beta}_{\zeta}],\quad Z_{\zeta}=|C_{\zeta}|,

where 𝜷ζ=𝜷ap+aζ.\boldsymbol{\beta}_{\zeta}=\boldsymbol{\beta}-a\cdot p+a\cdot\zeta. The following is the second half of Theorem 1.4.

Proposition 8.2 (BPZ-Cardy equations in 𝔻\mathbb{D}).

In the (𝔻,0)(\mathbb{D},0)-uniformization we have

2a2(12θ2+(θlogZζ)θ)𝐄^ζX=ˇvζ𝐄^ζX,(ζ=eiθ,θ),-\frac{2}{a^{2}}\Big{(}\frac{1}{2}\partial_{\theta}^{2}+\big{(}\partial_{\theta}\log Z_{\zeta}\big{)}\partial_{\theta}\Big{)}\widehat{\mathbf{E}}_{\zeta}X=\check{\mathcal{L}}_{v_{\zeta}}\widehat{\mathbf{E}}_{\zeta}X,\qquad(\zeta=\mathrm{e}^{i\theta},\theta\in\mathbb{R}),

where the Lie derivative ˇvζ\check{\mathcal{L}}_{v_{\zeta}} is taken over the finite notes of XX and supp𝛃ζ{ζ}.\mathrm{supp}\,\boldsymbol{\beta}_{\zeta}\setminus\{\zeta\}.

Proof.

Let ζ=eiθD.\zeta=\mathrm{e}^{i\theta}\in\partial D. Applying the BPZ equation in (𝔻,0)(\mathbb{D},0) to (Λζs)effX,(\Lambda^{s}_{\zeta})^{\mathrm{eff}}X, we have

𝐄L(ζ)(Λζs)effX=ˇvζ𝐄(Λζs)effX,\mathbf{E}\,L(\zeta)(\Lambda^{s}_{\zeta})^{\mathrm{eff}}X=\check{\mathcal{L}}_{v_{\zeta}}\mathbf{E}\,(\Lambda^{s}_{\zeta})^{\mathrm{eff}}X,

where LL is the differential operator in Subsection 7.13:

L(z)=κ2z2z2+3zz+h, or L(eiθ)=κ(12θ2+ihθ)+h.L(z)=\frac{\kappa}{2}z^{2}\partial_{z}^{2}+3z\partial_{z}+h,\textrm{ or }L(\mathrm{e}^{i\theta})=-\kappa\big{(}\frac{1}{2}\partial_{\theta}^{2}+ih\partial_{\theta}\big{)}+h.

In particular, Cζ=𝐄(Λζs)effC_{\zeta}=\mathbf{E}\,(\Lambda^{s}_{\zeta})^{\mathrm{eff}} satisfies the null vector equation

(8.4) L(ζ)Cζ=ˇvζCζ.L(\zeta)C_{\zeta}=\check{\mathcal{L}}_{v_{\zeta}}C_{\zeta}.

By differentiation,

𝐄L(ζ)(Λζs)effX𝐄(Λζs)eff=(L(ζ)h)𝐄^ζX+κζ2ζCζCζζ𝐄^ζX+L(ζ)CζCζ𝐄^ζX.\frac{\mathbf{E}\,L(\zeta)(\Lambda^{s}_{\zeta})^{\mathrm{eff}}X}{\mathbf{E}\,(\Lambda^{s}_{\zeta})^{\mathrm{eff}}}=(L(\zeta)-h)\widehat{\mathbf{E}}_{\zeta}X+\kappa\,\zeta^{2}\frac{\partial_{\zeta}C_{\zeta}}{C_{\zeta}}\partial_{\zeta}\widehat{\mathbf{E}}_{\zeta}X+\frac{L(\zeta)C_{\zeta}}{C_{\zeta}}\widehat{\mathbf{E}}_{\zeta}X.

It follows from Leibniz rule for Lie derivative that

ˇvζ𝐄(Λζs)effX𝐄(Λζs)eff=ˇvζ𝐄^ζX+ˇvζCζCζ𝐄^ζX.\frac{\check{\mathcal{L}}_{v_{\zeta}}\mathbf{E}\,(\Lambda^{s}_{\zeta})^{\mathrm{eff}}X}{\mathbf{E}\,(\Lambda^{s}_{\zeta})^{\mathrm{eff}}}=\check{\mathcal{L}}_{v_{\zeta}}\widehat{\mathbf{E}}_{\zeta}X+\frac{\check{\mathcal{L}}_{v_{\zeta}}C_{\zeta}}{C_{\zeta}}\,\widehat{\mathbf{E}}_{\zeta}X.

Combining all of the above, we have

(8.5) κ(12θ2+i(h+ζζCζCζ)θ)𝐄^ζX=ˇvζ𝐄^ζX-\kappa\Big{(}\frac{1}{2}\partial_{\theta}^{2}+i\big{(}h+\zeta\frac{\partial_{\zeta}C_{\zeta}}{C_{\zeta}}\big{)}\partial_{\theta}\Big{)}\widehat{\mathbf{E}}_{\zeta}X=\check{\mathcal{L}}_{v_{\zeta}}\widehat{\mathbf{E}}_{\zeta}X

in the (𝔻,0)(\mathbb{D},0)-uniformization. Note that the coefficient of 𝐄^ζX\widehat{\mathbf{E}}_{\zeta}X vanishes since CζC_{\zeta} satisfies the null vector equation (8.4).

Let

𝜷ζ0=aζ+(ba2)q+(ba2)q,Cζ0=C(b)[𝜷ζ0].\boldsymbol{\beta}_{\zeta}^{0}=a\cdot\zeta+\Big{(}b-\frac{a}{2}\Big{)}\cdot q+\Big{(}b-\frac{a}{2}\Big{)}\cdot q^{*},\qquad C_{\zeta}^{0}=C_{(b)}[\boldsymbol{\beta}_{\zeta}^{0}].

Then we have

ζζCζ0Cζ0=h,h=h1,2:=12a2ab.\zeta\frac{\partial_{\zeta}C_{\zeta}^{0}}{C_{\zeta}^{0}}=-h,\quad h=h_{1,2}:=\frac{1}{2}a^{2}-ab.

and rewrite (8.5) as

κ(12θ2+(θlogCζCζ0)θ)𝐄^ζX=ˇvζ𝐄^ζX.-\kappa\Big{(}\frac{1}{2}\partial_{\theta}^{2}+\big{(}\partial_{\theta}\log\frac{C_{\zeta}}{C^{0}_{\zeta}}\big{)}\partial_{\theta}\Big{)}\widehat{\mathbf{E}}_{\zeta}X=\check{\mathcal{L}}_{v_{\zeta}}\widehat{\mathbf{E}}_{\zeta}X.

It is easy to see that

θlogZζ=θlogCζCζ0\partial_{\theta}\log Z_{\zeta}=\partial_{\theta}\log\frac{C_{\zeta}}{C^{0}_{\zeta}}

in the 𝔻\mathbb{D}-uniformization. ∎

8.3. Martingale-observables for chordal/radial SLE with forces and spins

We now prove Theorem 1.5.

Proof of Theorem 1.5.

We first consider the chordal case. Let gtg_{t} be the chordal SLE[𝜷]\mathrm{SLE}[\boldsymbol{\beta}] map driven by the real process ξt:\xi_{t}:

dξt=κdBt+λ(t)dt,λ(t)=(λgt1),λ=κξlogZξ,\mathrm{d}\xi_{t}=\sqrt{\kappa}\,\mathrm{d}B_{t}+\lambda(t)\,\mathrm{d}t,\quad\lambda(t)=(\lambda\,\|\,g_{t}^{-1}),\quad\lambda=\kappa\,\partial_{\xi}\log Z_{\xi},

Then we have

Mt=m(ξt,t),m(ξ,t)=(Rξgt1).M_{t}=m(\xi_{t},t),\quad m(\xi,t)=(R_{\xi}\,\|\,g_{t}^{-1}).

It follows from Itô’s formula that

dMt\displaystyle\mathrm{d}M_{t} =κξ|ξ=ξtm(ξ,t)dBt+κξ|ξ=ξtm(ξ,t)ξ|ξ=ξt(logZξgt1)dt\displaystyle=\sqrt{\kappa}\,\partial_{\xi}\big{|}_{\xi=\xi_{t}}m(\xi,t)\,\mathrm{d}B_{t}+\kappa\,\partial_{\xi}\big{|}_{\xi=\xi_{t}}m(\xi,t)\,\partial_{\xi}\big{|}_{\xi=\xi_{t}}(\log Z_{\xi}\,\|\,g_{t}^{-1})\,\mathrm{d}t
+κ2ξ2|ξ=ξtm(ξ,t)dt+Ltdt,\displaystyle+\frac{\kappa}{2}\,\partial_{\xi}^{2}\big{|}_{\xi=\xi_{t}}m(\xi,t)\,\mathrm{d}t+L_{t}\,\mathrm{d}t,

where

Lt:=dds|s=0(Rξtgt+s1).L_{t}:=\frac{\mathrm{d}}{\mathrm{d}s}\Big{|}_{s=0}\big{(}R_{\xi_{t}}\,\|\,g_{t+s}^{-1}\big{)}.

At each time t,t, fs,t:=gt+sgt1f_{s,t}:=g_{t+s}\circ g_{t}^{-1} satisfy the differential equations

ddsfs,t(ζ)=2fs,t(ζ)ξt+s.\frac{\mathrm{d}}{\mathrm{d}s}f_{s,t}(\zeta)=\frac{2}{f_{s,t}(\zeta)-\xi_{t+s}}.

As s0,s\to 0,

fs,t=id2skξt+o(s),kξ(z)=1ξz.f_{s,t}=\mathrm{id}-2sk_{\xi_{t}}+o(s),\qquad k_{\xi}(z)=\frac{1}{\xi-z}.

In terms of the time-dependent flows fs,t,f_{s,t}, LtL_{t} is rewritten as

Lt=dds|s=0(Rξtgt1fs,t1)=2(ˇkξtRξtgt1).L_{t}=\frac{\mathrm{d}}{\mathrm{d}s}\Big{|}_{s=0}\big{(}R_{\xi_{t}}\,\|\,g_{t}^{-1}\circ f_{s,t}^{-1}\big{)}=-2\big{(}\check{\mathcal{L}}_{k_{\xi_{t}}}R_{\xi_{t}}\,\|\,g_{t}^{-1}\big{)}.

The last equality follows from the fact any fields in 𝜷\mathcal{F}_{\boldsymbol{\beta}} depend smoothly on local charts. Thus the drift term of dMt\mathrm{d}M_{t} simplifies to

(κ2ξ2|ξ=ξtm(ξ,t)+κξ|ξ=ξtm(ξ,t)ξ|ξ=ξt(Zξgt1)(Zξgt1)2(ˇkξtRξtgt1))dt.\Big{(}\frac{\kappa}{2}\,\partial_{\xi}^{2}\big{|}_{\xi=\xi_{t}}m(\xi,t)+\kappa\,\partial_{\xi}\big{|}_{\xi=\xi_{t}}m(\xi,t)\frac{\partial_{\xi}|_{\xi=\xi_{t}}(Z_{\xi}\,\|\,g_{t}^{-1})}{(Z_{\xi}\,\|\,g_{t}^{-1})}-2\big{(}\check{\mathcal{L}}_{k_{\xi_{t}}}R_{\xi_{t}}\,\|\,g_{t}^{-1}\big{)}\Big{)}\,\mathrm{d}t.

It vanishes by the BPZ-Cardy equations (Proposition 8.1).

Next, we consider the radial case. For 𝒳=X1(z1)Xn(zn),Xj𝜷ˇ(D),\mathcal{X}=X_{1}(z_{1})\cdots X_{n}(z_{n}),X_{j}\in\mathcal{F}_{\check{\boldsymbol{\beta}\,}}(D), denote

Rζ(z1,,zn)𝐄^ζ[X1(z1)Xn(zn)].R_{\zeta}(z_{1},\cdots,z_{n})\equiv\widehat{\mathbf{E}}_{\zeta}[X_{1}(z_{1})\cdots X_{n}(z_{n})].

Then the process Mt(z1,,zn)M_{t}(z_{1},\cdots,z_{n}) is represented by

Mt=m(ζt,t),m(ζ,t)=(Rζgt1),M_{t}=m(\zeta_{t},t),\qquad m(\zeta,t)=\big{(}R_{\zeta}\,\|\,g_{t}^{-1}\big{)},

where gt:(Dt,γt,q)(𝔻,ζt,0)g_{t}:(D_{t},\gamma_{t},q)\to(\mathbb{D},\zeta_{t},0) is the radial SLE[𝜷]\mathrm{SLE}[\boldsymbol{\beta}] map driven by the real process θt:\theta_{t}:

dθt=κdBt+λ(t)dt,λ(t)=(λgt1),λ=κθlogZ𝜷ζ,ζt=eiθt.\mathrm{d}\theta_{t}=\sqrt{\kappa}\,\mathrm{d}B_{t}+\lambda(t)\,\mathrm{d}t,\quad\lambda(t)=(\lambda\,\|\,g_{t}^{-1}),\quad\lambda=\kappa\,\partial_{\theta}\log Z_{\boldsymbol{\beta}_{\zeta}},\quad\zeta_{t}=\mathrm{e}^{i\theta_{t}}.

Using a similar argument in the chordal case, we find the drift term of dMt\mathrm{d}M_{t} as

(κ2θ2+κθ(Zζgt1)(Zζgt1))θ)|ζ=ζtm(ζ,t)dt+(ˇvζtRζtgt1)dt.\Big{(}\frac{\kappa}{2}\partial_{\theta}^{2}+\kappa\frac{\partial_{\theta}(Z_{\zeta}\,\|\,g_{t}^{-1})}{(Z_{\zeta}\,\|\,g_{t}^{-1})}\big{)}\partial_{\theta}\Big{)}\Big{|}_{\zeta=\zeta_{t}}~{}m(\zeta,t)\,\mathrm{d}t\\ +\ \big{(}\check{\mathcal{L}}_{v_{\zeta_{t}}}R_{\zeta_{t}}\,\|\,g_{t}^{-1}\big{)}\mathrm{d}t.

By the BPZ-Cardy equations (Proposition 8.2) in the radial case, MtM_{t} is driftless. ∎

Example.

Non-chiral vertex fields

𝒱𝜷(σ):=eiσΦ𝜷=𝒪𝜷[σz,σz]\mathcal{V}_{\boldsymbol{\beta}}^{(\sigma)}:=\mathrm{e}^{*i\sigma\Phi_{\boldsymbol{\beta}}}=\mathcal{O}_{\boldsymbol{\beta}}[\sigma\cdot z,-\sigma\cdot z]

have real conformal dimensions [λ+,λ][\lambda^{+},\lambda^{-}] at z:z:

λ+=σ22σb,λ=σ22+σb.\lambda^{+}=\frac{\sigma^{2}}{2}-\sigma b,\qquad\lambda^{-}=\frac{\sigma^{2}}{2}+\sigma b.

If the conformal spin s:=λ+λs:=\lambda^{+}-\lambda^{-} is 1, then 𝒱𝜷(σ)\mathcal{V}_{\boldsymbol{\beta}}^{(\sigma)} transforms as a (formal) vector field and its flow lines can be identified with SLE[𝜷],\mathrm{SLE}[\boldsymbol{\beta}], see [26] for a chordal case and [27] for a radial case with spin at q.q.

8.4. Examples of radial SLE martingale-observables

In this subsection we present examples of radial SLE martingale-observables including Schramm-Sheffield’s observables, Friedrich-Werner’s formula, and the restriction formula in the standard radial case with the following background charge 𝜷\boldsymbol{\beta}:

(8.6) 𝜷=ap+(ba2)q+(ba2)q,\boldsymbol{\beta}=a\cdot p+\Big{(}b-\frac{a}{2}\Big{)}\cdot q+\Big{(}b-\frac{a}{2}\Big{)}\cdot q^{*},
Example (Schramm-Sheffield’s observables).

In the standard chordal case with

𝜷=ap+(2ba)q,\boldsymbol{\beta}=a\cdot p+(2b-a)\cdot q,

the 1-point functions of the bosonic fields

φ𝜷(z)=𝐄[Φ𝜷(z)]=2aargw(z)2bargw(z),w:(D,p,q)(,0,)\varphi_{\boldsymbol{\beta}}(z)=\mathbf{E}[\Phi_{\boldsymbol{\beta}}(z)]=2a\arg w(z)-2b\arg w^{\prime}(z),\quad w:(D,p,q)\to(\mathbb{H},0,\infty)

were introduced as SLE martingale-observables by Schramm and Sheffield, see [32]. Similarly, the 1-point functions φ𝜷=𝐄Φ𝜷\varphi_{\boldsymbol{\beta}}=\mathbf{E}\,\Phi_{\boldsymbol{\beta}} (with standard radial background charge 𝜷\boldsymbol{\beta} in (8.6)) of the bosonic fields are martingale-observables of radial SLEs:

φ𝜷(z)=𝐄[Φ𝜷(z)]=2aarg(1w(z))aargw(z)2bargw(z)w(z),\varphi_{\boldsymbol{\beta}}(z)=\mathbf{E}[\Phi_{\boldsymbol{\beta}}(z)]=2a\arg(1-w(z))-a\arg w(z)-2b\arg\frac{w^{\prime}(z)}{w(z)},

where w:(D,p,q)(𝔻,1,0)w:(D,p,q)\to(\mathbb{D},1,0) is a conformal map. By Itô’s calculus, we have

φt(z)\displaystyle\varphi_{t}(z) =𝐄[ΦDt,γt,q(z)]=2aarg(1wt(z))aargwt(z)2bargwt(z)wt(z)\displaystyle=\mathbf{E}[\Phi_{D_{t},\gamma_{t},q}(z)]=2a\arg(1-w_{t}(z))-a\arg w_{t}(z)-2b\arg\frac{w_{t}^{\prime}(z)}{w_{t}(z)}
=2aarg(1w(z))aargw(z)2bargw(z)w(z)+20tRe1+ws(z)1ws(z)dBs.\displaystyle=2a\arg(1-w(z))-a\arg w(z)-2b\arg\frac{w^{\prime}(z)}{w(z)}+\sqrt{2}\int_{0}^{t}\mathrm{Re}\,\frac{1+w_{s}(z)}{1-w_{s}(z)}\,\mathrm{d}B_{s}.

One can use the 22-point martingale-observables

𝐄[Φ𝜷(z1)Φ𝜷(z2)]=2G(z1,z2)+φ𝜷(z1)φ𝜷(z2)\mathbf{E}[\Phi_{\boldsymbol{\beta}}(z_{1})\Phi_{\boldsymbol{\beta}}(z_{2})]=2G(z_{1},z_{2})+\varphi_{\boldsymbol{\beta}}(z_{1})\varphi_{\boldsymbol{\beta}}(z_{2})

or Hadamard’s variation formula

(8.7) dGDt(z1,z2)=Re1+wt(z1)1wt(z1)Re1+wt(z2)1wt(z2)dt=12dφ𝜷(z1),φ𝜷(z2)t\mathrm{d}G_{D_{t}}(z_{1},z_{2})=-\,\mathrm{Re}\,\frac{1+w_{t}(z_{1})}{1-w_{t}(z_{1})}\,\mathrm{Re}\,\frac{1+w_{t}(z_{2})}{1-w_{t}(z_{2})}\,\mathrm{d}t=-\frac{1}{2}\,\mathrm{d}\langle\varphi_{\boldsymbol{\beta}}(z_{1}),\varphi_{\boldsymbol{\beta}}(z_{2})\rangle_{t}

to construct a coupling of radial SLE and the Gaussian free field such that

𝐄[ΦD,p,q|γ[0,t]]=ΦDt,γt,q,\mathbf{E}[\,\Phi_{D,p,q}\,|\,\gamma[0,t]\,]=\Phi_{D_{t},\gamma_{t},q},

see [14].

Let us recall the restriction property of radial SLE(8/3)\mathrm{SLE}(8/3) (see [21, Section 6.5]):

  • the law of SLE(8/3)\mathrm{SLE}(8/3) in 𝔻\mathbb{D} conditioned to avoid a fixed hull KK is identical to the law of SLE(8/3)\mathrm{SLE}(8/3) in 𝔻K;\mathbb{D}\setminus K;

  • equivalently, there exist λ\lambda and μ\mu such that for all K,K,

    𝐏(SLE(8/3) avoids K)=|ΨK(1)|λ(ΨK(0))μ,\mathbf{P}(\mathrm{SLE}(8/3)\textrm{ avoids }K)=|\Psi_{K}^{\prime}(1)|^{\lambda}(\Psi_{K}^{\prime}(0))^{\mu},

    where ΨK\Psi_{K} is the conformal map (𝔻K,0)(𝔻,0)(\mathbb{D}\setminus K,0)\to(\mathbb{D},0) satisfying ΨK(0)>0.\Psi_{K}^{\prime}(0)>0. (The restriction exponents λ\lambda and μ\mu of radial SLE(8/3)\mathrm{SLE}(8/3) are equal to 5/85/8 and 5/48,5/48, respectively.)

Let κ4.\kappa\leq 4. On the event γ[0,)K=,\gamma[0,\infty)\cap K=\emptyset, a conformal map ht:Ωt=gt(DtK)𝔻h_{t}:\Omega_{t}=g_{t}(D_{t}\setminus K)\to\mathbb{D} is defined by

ht=g~tΨKgt1,h_{t}=\widetilde{g}_{t}\circ\Psi_{K}\circ g_{t}^{-1},

where g~t\widetilde{g}_{t} is a Loewner map from D~t=Dγ~[0,t]\widetilde{D}_{t}=D\setminus\widetilde{\gamma}[0,t] onto 𝔻,\mathbb{D}, and γ~(t)=ΨKγ(t).\widetilde{\gamma}(t)=\Psi_{K}\circ\gamma(t). Let Λeff\Lambda^{\mathrm{eff}} be the effective one-leg operator, i.e., Λeff=𝒫qΛ\Lambda^{\mathrm{eff}}=\mathcal{P}_{q}\Lambda and

Mt:=(Z𝜷tidΩt)(Z𝜷tid),𝜷t=aζt+(ba+β2)q+(ba+β2)q,(ζt=eiκBt).M_{t}:=\frac{(Z_{\boldsymbol{\beta}_{t}}\|\,\mathrm{id}_{\Omega_{t}})}{(Z_{\boldsymbol{\beta}_{t}}\|\,\mathrm{id}_{\mathbb{H}})},\quad\boldsymbol{\beta}_{t}=a\cdot\zeta_{t}+(b-\frac{a+\beta}{2})\cdot q+(b-\frac{a+\beta}{2})\cdot q^{*},\qquad(\zeta_{t}=\mathrm{e}^{i\sqrt{\kappa}B_{t}}).

Then

Mt=|ht(ζt)|λht(0)μ=(ζtht(ζt)ht(ζt))λht(0)μ,M_{t}=|h_{t}^{\prime}(\zeta_{t})|^{\lambda}h_{t}^{\prime}(0)^{\mu}=\Big{(}\zeta_{t}\frac{h_{t}^{\prime}(\zeta_{t})}{h_{t}(\zeta_{t})}\Big{)}^{\lambda}h_{t}^{\prime}(0)^{\mu},

where exponents are given by

λ=λb(Λeff)a22ab=6κ2κ,μ=Hq(Λeff)a24b2=(κ2)(6κ)8κ.\lambda=\lambda_{b}(\Lambda^{\mathrm{eff}})\equiv\frac{a^{2}}{2}-ab=\frac{6-\kappa}{2\kappa},\qquad\mu=H_{q}(\Lambda^{\mathrm{eff}})\equiv\frac{a^{2}}{4}-b^{2}=\frac{(\kappa-2)(6-\kappa)}{8\kappa}.

Restriction property of radial SLE(8/3)\mathrm{SLE}(8/3) follows from the local martingale property of MtM_{t} (by optional stopping theorem). This is a special case of the following formula:

(8.8) the drift term of dMt=c6ζt2Sht(ζt)Mtdt.\textrm{the drift term of }\mathrm{d}M_{t}=-\frac{c}{6}\zeta_{t}^{2}S_{h_{t}}(\zeta_{t})M_{t}\,\mathrm{d}t.

In Subsection 8.6 we use the CFT argument to prove (8.8) for radial SLE[𝜷].\mathrm{SLE}[\boldsymbol{\beta}].

We now prove Friedrich-Werner’s formula in the radial case.

Theorem 8.3.

Let θj(0,2π)\theta_{j}\in(0,2\pi) all distinct. Then we have

limt01tn𝐏(SLE(8/3) hits all [rteiθj,eiθj])=(2)ne2ij=1nθj𝐄[T𝜷(eiθ1)T𝜷(eiθn)id𝔻],\lim_{t\to 0}\frac{1}{t^{n}}\mathbf{P}(\mathrm{SLE}(8/3)\textrm{ hits all }[r_{t}\mathrm{e}^{i\theta_{j}},\mathrm{e}^{i\theta_{j}}])=(-2)^{n}\mathrm{e}^{2i\sum_{j=1}^{n}\theta_{j}}\mathbf{E}\,[\,T_{\boldsymbol{\beta}}(\mathrm{e}^{i\theta_{1}})\cdots T_{\boldsymbol{\beta}}(\mathrm{e}^{i\theta_{n}})\,\|\,\mathrm{id}_{\mathbb{D}}\,],

where 𝛃=ap+(ba/2)q+(ba/2)q,\boldsymbol{\beta}=a\cdot p+(b-a/2)\cdot q+(b-a/2)\cdot q^{*}, a=123,a=\frac{1}{2}\sqrt{3}, b=163,b=-\frac{1}{6}\sqrt{3}, p=1,q=0,p=1,q=0, and rt=12t.r_{t}=1-2\sqrt{t}.

Proof.

Let zj=eiθj.z_{j}=\mathrm{e}^{i\theta_{j}}. We apply Ward’s equations to the function

𝐄[T𝜷(z)T𝜷(z1)T𝜷(zn)id𝔻]=𝐄[T𝜷ˇ(z)Λeff(1)T𝜷ˇ(z1)T𝜷ˇ(zn)id𝔻],\mathbf{E}\,[\,T_{\boldsymbol{\beta}}(z)T_{\boldsymbol{\beta}}(z_{1})\cdots T_{\boldsymbol{\beta}}(z_{n})\,\|\,\mathrm{id}_{\mathbb{D}}\,]=\mathbf{E}\,[\,T_{\check{\boldsymbol{\beta}\,}}(z)\Lambda^{\mathrm{eff}}(1)T_{\check{\boldsymbol{\beta}\,}}(z_{1})\cdots T_{\check{\boldsymbol{\beta}\,}}(z_{n})\,\|\,\mathrm{id}_{\mathbb{D}}\,],

by replacing T𝜷ˇ(z)T_{\check{\boldsymbol{\beta}\,}}(z) on the right-hand side with the corresponding Ward’s functional. Denote 𝒛=(z1,,zn),\boldsymbol{z}=(z_{1},\cdots,z_{n}), 𝒛j=(z1,,z^j,zn),\boldsymbol{z}_{j}=(z_{1},\cdots,\widehat{z}_{j}\cdots,z_{n}), and

R(ζ;𝒛)=𝐄[Λeff(ζ)T𝜷ˇ(z1)T𝜷ˇ(zn)].R(\zeta;\boldsymbol{z})=\mathbf{E}\,[\,\Lambda^{\mathrm{eff}}(\zeta)\,T_{\check{\boldsymbol{\beta}\,}}(z_{1})\cdots T_{\check{\boldsymbol{\beta}\,}}(z_{n})\,].

The non-random field R(ζ;𝒛)R(\zeta;\boldsymbol{z}) is a boundary differential of conformal dimension λ=12a2ab=5/8\lambda=\frac{1}{2}{a^{2}}-ab=5/8 with respect to ζ\zeta and of conformal dimension 22 with respect to zj.z_{j}. It is also a differential of conformal dimension μ=14a2b2=5/48\mu=\frac{1}{4}a^{2}-b^{2}=5/48 with respect to q=0.q=0. It follows from Ward’s equation for Λeff\Lambda^{\mathrm{eff}} that

(8.9) R(ζ;z,𝒛)=12z2(vz,𝔻)R(ζ;𝒛),(in id𝔻),R(\zeta;z,\boldsymbol{z})=\frac{1}{2z^{2}}\,\mathcal{L}({v_{z}},\mathbb{D})\,R(\zeta;\boldsymbol{z}),\quad(\textrm{in }\mathrm{id}_{\mathbb{D}}),

where vz(η)=η(z+η)/(zη).v_{z}(\eta)=\eta(z+\eta)/(z-\eta).

Let

U(θ1,,θn)=limt0(1)n(2t)ne2ij=1nθj𝐏(SLE(8/3) hits all [rteiθj,eiθj])U(\theta_{1},\cdots,\theta_{n})=\lim_{t\to 0}\frac{(-1)^{n}}{(2t)^{n}}\mathrm{e}^{-2i\sum_{j=1}^{n}\theta_{j}}\mathbf{P}(\mathrm{SLE}(8/3)\textrm{ hits all }[r_{t}\mathrm{e}^{i\theta_{j}},\mathrm{e}^{i\theta_{j}}])

(if the limit exists). Define the non-random field T(ζ;z1,,zn)T(\zeta;z_{1},\cdots,z_{n}) as follows:

  • TT is a boundary differential of conformal dimension λ=5/8\lambda=5/8 with respect to ζ,\zeta, and of conformal dimension 22 with respect to zj;z_{j};

  • TT is a differential of conformal dimension μ=14a2b2=5/48\mu=\frac{1}{4}a^{2}-b^{2}=5/48 with respect to q=0;q=0;

  • (T(eiφ;eiθ1,,eiθn)id𝔻)=U(θ1φ,,θnφ).(T(\mathrm{e}^{i\varphi};\mathrm{e}^{i\theta_{1}},\cdots,\mathrm{e}^{i\theta_{n}})\,\|\,\mathrm{id}_{\mathbb{D}})=U(\theta_{1}-\varphi,\cdots,\theta_{n}-\varphi).

We now claim that if the limit U(θ1,,θn)U(\theta_{1},\cdots,\theta_{n}) exists then the limit U(θ,θ1,,θn)U(\theta,\theta_{1},\cdots,\theta_{n}) exists and

(8.10) T(1;z,𝒛)=12z2(vz,𝔻)T(1;𝒛)z,zj𝔻.T(1;z,\boldsymbol{z})=\frac{1}{2z^{2}}\,\mathcal{L}({v_{z}},\mathbb{D})\,T(1;\boldsymbol{z})\qquad z,z_{j}\in\partial\mathbb{D}.

By (8.9) and (8.10), T(1;)T(1;\cdot) and R(1;)R(1;\cdot) satisfy the same recursive equation (see the remark at the end of this subsection) and are therefore equal since T(1;)=R(1;)=1T(1;\cdot)=R(1;\cdot)=1 for n=0.n=0. Thus

U(θ1,,θn)=R(1;eiθ1,,eiθn).U(\theta_{1},\cdots,\theta_{n})=R(1;\mathrm{e}^{i\theta_{1}},\cdots,\mathrm{e}^{i\theta_{n}}).

To verify the induction argument for existence of the limit U(θ1,,θn)U(\theta_{1},\cdots,\theta_{n}) and show (8.10), denote 𝜽=(θ1,,θn).\boldsymbol{\theta}=(\theta_{1},\cdots,\theta_{n}). We write 𝐏(𝜽)\mathbf{P}(\boldsymbol{\theta}) for the probability that radial SLE(8/3)\mathrm{SLE}(8/3) path hits all segments [rteiθj,eiθj](1jn)[r_{t}\mathrm{e}^{i\theta_{j}},\mathrm{e}^{i\theta_{j}}]\,(1\leq j\leq n) and 𝐏(𝜽|¬θ)\mathbf{P}(\boldsymbol{\theta}\,|\,\neg\theta) for the same probability conditioned on the event that the path avoids [rteiθ,eiθ].[r_{t}\mathrm{e}^{i\theta},\mathrm{e}^{i\theta}]. By the induction hypothesis,

(8.11) 𝐏(𝜽)(2t)nz12zn2T(1;𝒛),\mathbf{P}(\boldsymbol{\theta})\approx(-2t)^{n}z_{1}^{2}\cdots z_{n}^{2}\,T(1;\boldsymbol{z}),

where zj=eiθj.z_{j}=\mathrm{e}^{i\theta_{j}}. On the other hand, by the restriction property of radial SLE(8/3),\mathrm{SLE}(8/3), we have

(8.12) 1𝐏(θ)=|ψt(1)|λψt(0)16λ1-\mathbf{P}(\theta)=|\psi_{t}^{\prime}(1)|^{\lambda}\,\psi_{t}^{\prime}(0)^{\frac{1}{6}\lambda}

and

(8.13) 𝐏(𝜽|¬θ)(2t)nT(ψt(1);ψt(z1),,ψt(zn))j=1nzj2ψt(zj)2,\mathbf{P}(\boldsymbol{\theta}\,|\,\neg\theta)\approx(-2t)^{n}\,T(\psi_{t}(1);\psi_{t}(z_{1}),\cdots,\psi_{t}(z_{n}))\prod_{j=1}^{n}z_{j}^{2}\psi_{t}^{\prime}(z_{j})^{2},

where ψt\psi_{t} is a slit map from (𝔻[rteiθ,eiθ],0)(\mathbb{D}\setminus[r_{t}\mathrm{e}^{i\theta},\mathrm{e}^{i\theta}],0) onto (𝔻,0)(\mathbb{D},0) with ψt(0)>0\psi_{t}^{\prime}(0)>0. It follows from (8.11) – (8.13) and 𝐏(θ,𝜽)=𝐏(𝜽)𝐏(𝜽|¬θ)(1𝐏(θ))\mathbf{P}(\theta,\boldsymbol{\theta})=\mathbf{P}(\boldsymbol{\theta})-\mathbf{P}(\boldsymbol{\theta}\,|\,\neg\theta)(1-\mathbf{P}(\theta))(z=eiθ)(z=\mathrm{e}^{i\theta}) that

𝐏(θ,𝜽)z2z12zn2(2t)n+1=ψt(0)λ/6|ψt(1)|λj=1nψt(zj)2T(ψt(1);ψt(z1),,ψt(zn))T(1;𝒛)2tz2\frac{\mathbf{P}(\theta,\boldsymbol{\theta})}{z^{2}z_{1}^{2}\cdots z_{n}^{2}(-2t)^{n+1}}=\frac{\psi_{t}^{\prime}(0)^{\lambda/6}|\psi_{t}^{\prime}(1)|^{\lambda}\prod_{j=1}^{n}\psi_{t}^{\prime}(z_{j})^{2}T(\psi_{t}(1);\psi_{t}(z_{1}),\cdots,\psi_{t}(z_{n}))-T(1;\boldsymbol{z})}{2tz^{2}}

up to o(t)o(t) terms. Thus the limit U(θ,θ1,,θn)U(\theta,\theta_{1},\cdots,\theta_{n}) exists. Since ψt(0)=et,\psi_{t}^{\prime}(0)=\mathrm{e}^{t}, we have (8.10):

T(1;z,𝒛)=12z2((vz,𝔻{0})+λ6)T(1;𝒛)=12z2(vz,𝔻)T(1;𝒛).T(1;z,\boldsymbol{z})=\frac{1}{2z^{2}}\Big{(}\mathcal{L}(v_{z},\mathbb{D}\setminus\{0\})+\frac{\lambda}{6}\Big{)}T(1;\boldsymbol{z})=\frac{1}{2z^{2}}\,\mathcal{L}({v_{z}},\mathbb{D})\,T(1;\boldsymbol{z}).

Remark.

The formula (8.9) holds for all κ.\kappa. Setting R(z,𝒛)R(1;z,𝒛),R(z,\boldsymbol{z})\equiv R(1;z,\boldsymbol{z}), the formula (8.9) at ζ=1\zeta=1 gives the following recursive formula for R:R:

R(z,𝒛)\displaystyle R(z,\boldsymbol{z}) =12z2(2n1+z1z+2λz(1z)2+(a24b2)+1+z1zj=1nzjzj)R(𝒛)\displaystyle=\frac{1}{2z^{2}}\Big{(}2n\frac{1+z}{1-z}+2\lambda\frac{z}{(1-z)^{2}}+(\frac{a^{2}}{4}-b^{2})+\frac{1+z}{1-z}\sum_{j=1}^{n}z_{j}\partial_{z_{j}}\Big{)}R(\boldsymbol{z})
+12z2j=1n(zjz+zjzzjzj+2z2+2zzjzj2(zzj)2)R(𝒛)+c2j=1n1(zzj)4R(𝒛j).\displaystyle+\frac{1}{2z^{2}}\sum_{j=1}^{n}\Big{(}z_{j}\frac{z+z_{j}}{z-z_{j}}\partial_{z_{j}}+2\frac{z^{2}+2zz_{j}-z_{j}^{2}}{(z-z_{j})^{2}}\Big{)}R(\boldsymbol{z})+\frac{c}{2}\sum_{j=1}^{n}\frac{1}{(z-z_{j})^{4}}R(\boldsymbol{z}_{j}).

8.5. 1-point vertex observables

In this subsection we discuss some basic examples of 1-point vertex observables, including Lawler-Schramm-Werner’s derivative exponents of radial SLEs on the boundary. Let

M(τ+,τ;τq+,τq)(z)=𝐄^𝒪𝜷ˇ[𝝉]=𝐄𝒪𝜷[𝝉],𝝉=τ+z+τz+τq+q+τqq.M^{(\tau^{+},\tau^{-};\tau_{q}^{+},\tau_{q}^{-})}(z)=\widehat{\mathbf{E}}\,\mathcal{O}_{\check{\boldsymbol{\beta}\,}}[\boldsymbol{\tau}]=\mathbf{E}\,\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}],\qquad\boldsymbol{\tau}=\tau^{+}\cdot z+\tau^{-}\cdot z^{*}+\tau_{q}^{+}\cdot q+\tau_{q}^{-}\cdot q^{*}.

Then by (6.2) we have

M(τ+,τ;τq+,τq)=(wq)hq+(wq¯)hq(w)h+(w¯)hwν+(w¯)ν(1w)aτ+(1w¯)aτ(1|w|2)τ+τ,M^{(\tau^{+},\tau^{-};\tau_{q}^{+},\tau_{q}^{-})}=(w_{q}^{\prime})^{h_{q}^{+}}(\overline{w_{q}^{\prime}})^{h_{q}^{-}}(w^{\prime})^{h^{+}}(\overline{w^{\prime}})^{h^{-}}w^{\nu^{+}}(\bar{w})^{\nu^{-}}(1-w)^{a\tau^{+}}(1-\bar{w})^{a\tau^{-}}(1-|w|^{2})^{\tau^{+}\tau^{-}},

where the exponents are ν±=τ±(τq±+ba/2)\nu^{\pm}=\tau^{\pm}(\tau_{q}^{\pm}+b-a/2) and the dimensions are

h±=λb(τ±),hq±=τq±(τq±a)2.h^{\pm}=\lambda_{b}(\tau^{\pm}),\qquad h_{q}^{\pm}=\frac{\tau_{q}^{\pm}(\tau_{q}^{\pm}-a)}{2}.

The last formulas come from hq±=λb(τq±+ba/2)λb(τq±).h_{q}^{\pm}=\lambda_{b}(\tau_{q}^{\pm}+b-a/2)-\lambda_{b}(\tau_{q}^{\pm}).

• Constant fields.  The simplest examples of 1-point vertex fields are constant fields, i.e., vertex fields with τ±=0.\tau^{\pm}=0. By the neutrality condition, τq=τq+.\tau_{q}^{-}=-\tau_{q}^{+}. Since wt(0)=etiκBt,w_{t}^{\prime}(0)=\mathrm{e}^{t-i\sqrt{\kappa}B_{t}},

Mt(0,0;τq,τq)=eτq2t+iaκτqBt=eτq2t+i2τqBt.M_{t}^{(0,0;\tau_{q},-\tau_{q})}=\mathrm{e}^{\tau_{q}^{2}t+ia\sqrt{\kappa}\tau_{q}B_{t}}=\mathrm{e}^{\tau_{q}^{2}t+i\sqrt{2}\tau_{q}B_{t}}.

This is a martingale.

• Real fields.  The 1-point vertex fields are real if and only if τ+=τ\tau^{+}=\tau^{-} and τq+=τq.\tau_{q}^{+}=\tau_{q}^{-}. By the neutrality condition, τq+=τ+.\tau_{q}^{+}=-\tau^{+}. Thus the only real fields are

M(τ,τ;τ,τ)=|wq|τ2+aτ|w|τ22bτ|w|τ(2ba2τ)|1w|2aτ(1|w|2)τ2.M^{(\tau,\tau;-\tau,-\tau)}=|w_{q}^{\prime}|^{\tau^{2}+a\tau}|w^{\prime}|^{\tau^{2}-2b\tau}|w|^{\tau(2b-a-2\tau)}|1-w|^{2a\tau}(1-|w|^{2})^{\tau^{2}}.

When τ=a,\tau=-a, there is no covariance at q.q. In this special case,

M(a,a;a,a)=|ww|12/κ(1|w|2|1w|2)2/κ.M^{(-a,-a;a,a)}=\bigg{|}\frac{w^{\prime}}{w}\bigg{|}^{1-2/\kappa}\bigg{(}\frac{1-|w|^{2}}{|1-w|^{2}}\bigg{)}^{2/\kappa}.
Example.

If κ=2,\kappa=2, then M(a,a;a,a)M^{(-a,-a;a,a)} coincides with the Lawler-Schramm-Werner observable

M=1|w|2|1w|2=P𝔻(1,w)P𝔻(1,0)=PD(p,z)PD(p,q),M=\frac{1-|w|^{2}}{|1-w|^{2}}=\frac{P_{\mathbb{D}}(1,w)}{P_{\mathbb{D}}(1,0)}=\frac{P_{D}(p,z)}{P_{D}(p,q)},

where PDP_{D} is the Poisson kernel of a domain D.D. As mentioned in Subsection 1.2, this 1-point field is an important observable in the theory of LERW.

Example.

If κ=4,\kappa=4, then

M=|ww|1/2(1|w|2|1w|2)1/2M=\bigg{|}\frac{w^{\prime}}{w}\bigg{|}^{1/2}\bigg{(}\frac{1-|w|^{2}}{|1-w|^{2}}\bigg{)}^{1/2}

is Beffara’s type observable for radial SLE(4),\mathrm{SLE}(4), see [3]. In the chordal case, Beffara’s observables are real martingale-observables of conformal dimensions

h+=h=12κ16,hq=0h^{+}=h^{-}=\frac{1}{2}-\frac{\kappa}{16},\qquad h_{q}=0

with the estimate

𝐏(z,ε)ε118κM(z),\mathbf{P}(z,\varepsilon)\asymp\varepsilon^{1-\frac{1}{8}\kappa}~{}M(z),

where 𝐏(z,ε)\mathbf{P}(z,\varepsilon) is the probability that the SLE(κ)\mathrm{SLE}(\kappa) curve (κ<8\kappa<8) hits the disc at zz of size ε(1)\varepsilon(\ll 1) measured in a local chart ϕ.\phi. See [6]. In [3], Feynman-Kac formula is used to construct radial SLE martingale-observables with the desired dimensions.

• 1-point vertex fields without covariance at qq.  A 1-point vertex field MM has no covariance at qq if and only if

(τq+,τq)=(0,0),(a,0),(0,a),or(a,a).(\tau_{q}^{+},\tau_{q}^{-})=(0,0),\quad(a,0),\quad(0,a),\quad\textrm{or}\quad(a,a).

The first case is just the non-chiral vertex field M=M(τ,τ;0,0).M=M^{(\tau,-\tau;0,0)}.

When τ=0,\tau^{-}=0, the 1-point vertex observable

M(τ,0;τq+,τq)=(wq)hq+(wq¯)hq(w)hwν(1w)aτM^{(\tau,0;\tau_{q}^{+},\tau_{q}^{-})}=(w^{\prime}_{q})^{h_{q}^{+}}(\overline{w^{\prime}_{q}})^{h_{q}^{-}}(w^{\prime})^{h}w^{\nu}(1-w)^{a\tau}

is holomorphic. Recall the expression for the exponents and the dimensions

h=12τ2bτ,ν=τ(b12a+τq+),hq±=τq±(τq±a)2.h=\frac{1}{2}\tau^{2}-b\tau,\quad\nu=\tau(b-\frac{1}{2}a+\tau_{q}^{+}),\quad h_{q}^{\pm}=\frac{\tau_{q}^{\pm}(\tau_{q}^{\pm}-a)}{2}.

• Holomorphic 1-point fields without spin at qq.  A holomorphic 1-point field MM has no spin at qq if and only if hq+=hq.h_{q}^{+}=h_{q}^{-}. Equivalently,

τq+=τqorτq++τq=a.\tau_{q}^{+}=\tau_{q}^{-}\quad\textrm{or}\quad\tau_{q}^{+}+\tau_{q}^{-}=a.

Case 1. τq+=τq.\tau_{q}^{+}=\tau_{q}^{-}. By the neutrality condition, we have

𝝉=τzτ2qτ2q.\boldsymbol{\tau}=\tau\cdot z-\frac{\tau}{2}\cdot q-\frac{\tau}{2}\cdot q^{*}.

The holomorphic 1-point field 𝒪𝜷[𝝉]\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}] is a generalization of the one-leg operator Λz.\Lambda_{z}. In this case, its conformal dimensions are

h=τ22bτ,hq±=hq=τ28+aτ4.h=\frac{\tau^{2}}{2}-b\tau,\quad h_{q}^{\pm}=h_{q}=\frac{\tau^{2}}{8}+\frac{a\tau}{4}.

Thus

Mt=e2hqt(wtwt)12τ2bτ(wt(1wt)2)12aτ.M_{t}=\mathrm{e}^{2h_{q}t}\Big{(}\frac{w_{t}^{\prime}}{w_{t}}\Big{)}^{\frac{1}{2}\tau^{2}-b\tau}\Big{(}\frac{w_{t}}{(1-w_{t})^{2}}\Big{)}^{-\frac{1}{2}{a\tau}}.
Example (Derivative exponents on the boundary [24]).

On the unit circle, (up to constant)

Mt(eiθ)=e2hqt|wt(eiθ)|h(sin2θt2)12aτ,M_{t}(\mathrm{e}^{i\theta})=\mathrm{e}^{2h_{q}t}|w_{t}^{\prime}(\mathrm{e}^{i\theta})|^{h}\Big{(}\sin^{2}\frac{\theta_{t}}{2}\Big{)}^{\frac{1}{2}a\tau},

where wt(eiθ)=eiθt.w_{t}(\mathrm{e}^{i\theta})=\mathrm{e}^{i\theta_{t}}. Given h,h, the equation h=12τ2bτh=\frac{1}{2}\tau^{2}-b\tau is solved by

τ±=a4(κ4±(κ4)2+16κh).\tau_{\pm}=\frac{a}{4}\big{(}\kappa-4\pm\sqrt{(\kappa-4)^{2}+16\kappa h}\big{)}.

With the choice of τ=τ+,\tau=\tau_{+}, Lawler, Schramm, and Werner proved that Mt(eiθ)M_{t}(\mathrm{e}^{i\theta}) is a martingale. They applied the optional stopping theorem to Mt(eiθ)M_{t}(\mathrm{e}^{i\theta}) and used the estimate

𝐄[Mt(eiθ)]e2hqt𝐄[|wt(eiθ)|h𝟏{τeiθ>t}]\mathbf{E}[M_{t}(\mathrm{e}^{i\theta})]\asymp\mathrm{e}^{2h_{q}t}\,\mathbf{E}[|w^{\prime}_{t}(\mathrm{e}^{i\theta})|^{h}\mathbf{1}_{\{\tau_{\mathrm{e}^{i\theta}}>t\}}]

to derive the derivative exponents:

𝐄[|wt(eiθ)|h𝟏{τeiθ>t}]e2hqt(sin2θ2)12aτ.\mathbf{E}[|w^{\prime}_{t}(\mathrm{e}^{i\theta})|^{h}\mathbf{1}_{\{\tau_{\mathrm{e}^{i\theta}}>t\}}]\asymp\mathrm{e}^{-2h_{q}t}\Big{(}\sin^{2}\frac{\theta}{2}\Big{)}^{\frac{1}{2}a\tau}.

(Recall that τz\tau_{z} is the first time when a point zz is swallowed by the hull of SLE, see Subsection 1.2.) From the derivative exponent for κ=6,\kappa=6, they obtained the annulus crossing exponent for SLE6\mathrm{SLE}_{6} and combined it with other exponents to prove Mandelbrot’s conjecture that the Hausdorff dimension of the planar Brownian frontier is 4/34/3. See [23] and references therein.

Example.

The field MM is a scalar if τ=2b.\tau=2b. In this case,

Mt=et(κ4)/8(wt+1wt2)(κ4)/(2κ).M_{t}=\mathrm{e}^{t(\kappa-4)/{8}}\Big{(}w_{t}+\frac{1}{w_{t}}-2\Big{)}^{(\kappa-4)/(2\kappa)}.

Its derivative M~=M\widetilde{M}=\partial M has the conformal dimensions [1,0;hq,hq].[1,0;h_{q},h_{q}]. It is not a vertex observable. If M~=M(τ~+,τ~;τ~q+,τ~q)\widetilde{M}=M^{(\tilde{\tau}^{+},\tilde{\tau}^{-};\tilde{\tau}_{q}^{+},\tilde{\tau}_{q}^{-})} with h~+=λb(τ~+)=1,\tilde{h}^{+}=\lambda_{b}(\tilde{\tau}^{+})=1, then τ~+=2a\tilde{\tau}^{+}=-2a or τ~+=2a+2b.\tilde{\tau}^{+}=2a+2b. As we will see below, the holomorphic 1-differentials without spin at qq are not forms of M(2b,0,b,b).\partial M^{(2b,0,-b,-b)}. Unlike the chordal case (see [18, Proposition 15.2]), the holomorphic differential observables are not necessarily vertex observables.

Example.

If we take τ=2ba\tau=2b-a so that h=h1,2:=12a2abh=h_{1,2}:=\frac{1}{2}a^{2}-ab (there are only two values of τ\tau such that h=h1,2,h=h_{1,2}, namely τ=a,2ba\tau=a,2b-a), then

𝝉=(2ba)z+(b+a2)q+(b+a2)q,\boldsymbol{\tau}=(2b-a)\cdot z+(-b+\frac{a}{2})\cdot q+(-b+\frac{a}{2})\cdot q^{*},

and

h=6κ2κ=aτ2,hq=(2κ)(6κ)16κ=h0,1/2,ν=0.h=\frac{6-\kappa}{2\kappa}=-\frac{a\tau}{2},\quad h_{q}=\frac{(2-\kappa)(6-\kappa)}{16\kappa}=-h_{0,1/2},\quad\nu=0.

In this case, we have

Mt=e2hqt(wt(1wt)2)(6κ)/(2κ).M_{t}=\mathrm{e}^{2h_{q}t}\Big{(}\frac{w_{t}^{\prime}}{(1-w_{t})^{2}}\Big{)}^{(6-\kappa)/(2\kappa)}.

If κ=2,\kappa=2, then

M=w(1w)2M=\frac{w^{\prime}}{(1-w)^{2}}

and its anti-derivative is 1/(1w).1/(1-w). See the next example.

Case 2. τq++τq=a.\tau_{q}^{+}+\tau_{q}^{-}=a. Let τq=τq+.\tau_{q}=\tau_{q}^{+}. It follows from the neutrality condition that

𝝉=az+τqq+(aτq)q.\boldsymbol{\tau}=-a\cdot z+\tau_{q}\cdot q+(a-\tau_{q})\cdot q^{*}.

Thus we have

M=(w)(κ2)/(2κ)wa(τq+12ab)(1w)2/κ|wq|τq2aτq.M=(w^{\prime})^{(\kappa-2)/(2\kappa)}w^{a(-\tau_{q}+\frac{1}{2}a-b)}(1-w)^{-2/\kappa}|w_{q}^{\prime}|^{\tau_{q}^{2}-a\tau_{q}}.
Example.

The fields MM have no covariance at qq if and only if τq=0\tau_{q}=0 or a.a. In these cases, we have

M(a,0;a,0)=(ww)(κ2)/(2κ)(1w)2/κM^{(-a,0;a,0)}=\Big{(}\frac{w^{\prime}}{w}\Big{)}^{(\kappa-2)/(2\kappa)}(1-w)^{-2/\kappa}

and

M(a,0;0,a)=(w)(κ2)/(2κ)w(6κ)/(2κ)(1w)2/κ.M^{(-a,0;0,a)}=(w^{\prime})^{(\kappa-2)/(2\kappa)}w^{(6-\kappa)/(2\kappa)}(1-w)^{-2/\kappa}.

For example, if κ=6,\kappa=6, then

M(a,0;0,a)=(w1w)1/3,M^{(-a,0;0,a)}=\Big{(}\dfrac{w^{\prime}}{1-w}\Big{)}^{1/3},

and if κ=2,\kappa=2, then both M(a,0;a,0)M^{(-a,0;a,0)} and M(a,0;0,a)M^{(-a,0;0,a)} produce the same observable

M(a,0;a,0)=1+M(a,0;0,a)=11w.M^{(-a,0;a,0)}=1+M^{(-a,0;0,a)}=\frac{1}{1-w}.

• Holomorphic differentials without spin at qq.  Let τ=τ+,\tau=\tau^{+}, τ=0.\tau^{-}=0. A holomorphic 1-point field MM is a 1-differential with respect to zz if and only if τ=2a\tau=-2a or τ=2(a+b).\tau=2(a+b). Furthermore, if MM has no spin at q,q, then τq+=τq.\tau_{q}^{+}=\tau_{q}^{-}. (If τ=2a\tau=-2a or τ=2(a+b),\tau=2(a+b), then the other possibility τq++τq=a\tau_{q}^{+}+\tau_{q}^{-}=a never happens because of the neutrality condition.)

Case 1. τ=2a.\tau=-2a. By the neutrality condition, we have

𝝉=2az+aq+aq\boldsymbol{\tau}=-2a\cdot z+a\cdot q+a\cdot q^{*}

and

M=ww2/κ1(1w)4/κ.M=w^{\prime}w^{2/\kappa-1}(1-w)^{-4/\kappa}.
Example.

If κ=2,\kappa=2, then

M(2a,0;a,a)=w(1w)2M^{(-2a,0;a,a)}=\frac{w^{\prime}}{(1-w)^{2}}

and its anti-derivative is 1/(1w).1/(1-w). See the previous example.

Example.

If κ=4,\kappa=4, then M(2a,0;a,a)=w(1w)1w1/2M^{(-2a,0;a,a)}=w^{\prime}(1-w)^{-1}w^{-1/2} and its anti-derivative is

log(1+w)log(1w).\log(1+\sqrt{w})-\log(1-\sqrt{w}).

Its imaginary part is a bosonic observable for a twisted conformal field theory.

Case 2. τ=2(a+b).\tau=2(a+b). By the neutrality condition, we have

𝝉=(2a+2b)z(a+b)q(a+b)q\boldsymbol{\tau}=(2a+2b)\cdot z-(a+b)\cdot q-(a+b)\cdot q^{*}

and

Mt=et(κ+4)/8wtwt3/2(1wt).M_{t}=\mathrm{e}^{t(\kappa+4)/8}w_{t}^{\prime}w_{t}^{-3/2}(1-w_{t}).

8.6. Restriction formulas

Let κ4.\kappa\leq 4. We first consider the chordal SLE[𝜷]\mathrm{SLE}[\boldsymbol{\beta}] with

𝜷=ap+βkqk+(2baβk)q,p,q,qkD.\boldsymbol{\beta}=a\cdot p+\sum\beta_{k}\cdot q_{k}+(2b-a-\sum\beta_{k})\cdot q,\qquad p,q,q_{k}\in\partial D.

Fix a hull K.K. Let gtg_{t} be the chordal SLE[𝜷]\mathrm{SLE}[\boldsymbol{\beta}] map with the hull Kt.K_{t}. On the event KK=,K_{\infty}\cap K=\emptyset, a conformal map ht:Ωt=gt(DtK)h_{t}:\Omega_{t}=g_{t}(D_{t}\setminus K)\to\mathbb{H} is defined by

ht=g~tΨKgt1,h_{t}=\widetilde{g}_{t}\circ\Psi_{K}\circ g_{t}^{-1},

where g~t\widetilde{g}_{t} is a Loewner map from D~t=DK~t\widetilde{D}_{t}=D\setminus\widetilde{K}_{t} onto ,\mathbb{H}, K~t=ΨK(Kt),\widetilde{K}_{t}=\Psi_{K}(K_{t}), and ΨK\Psi_{K} is the conformal map (K,0,)(,0,)(\mathbb{H}\setminus K,0,\infty)\to(\mathbb{H},0,\infty) satisfying ΨK()=1.\Psi_{K}^{\prime}(\infty)=1. Let

Mt:=(Z𝜷tidΩt)(Z𝜷tid)=𝐄(ΛΩteffid)(ξt,𝒒(t))𝐄(Λeffid)(ξt,𝒒(t)),M_{t}:=\frac{(Z_{\boldsymbol{\beta}_{t}}\|\,\mathrm{id}_{\Omega_{t}})}{(Z_{\boldsymbol{\beta}_{t}}\|\,\mathrm{id}_{\mathbb{H}})}=\frac{\mathbf{E}\,(\Lambda^{\mathrm{eff}}_{\Omega_{t}}\|\,\mathrm{id})(\xi_{t},\boldsymbol{q}(t))}{\mathbf{E}\,(\Lambda^{\mathrm{eff}}_{\mathbb{H}}\|\,\mathrm{id})(\xi_{t},\boldsymbol{q}(t))},

where

𝜷t=aξt+βkqk(t),dqk(t)=2qk(t)ξtdt,qk(0)=qk\boldsymbol{\beta}_{t}=a\cdot\xi_{t}+\sum\beta_{k}\cdot q_{k}(t),\qquad\mathrm{d}q_{k}(t)=\frac{2}{q_{k}(t)-\xi_{t}}\,\mathrm{d}t,\qquad q_{k}(0)=q_{k}

and Λeff=𝒫𝜷Λ\Lambda^{\mathrm{eff}}=\mathcal{P}_{\boldsymbol{\beta}}\,\Lambda is the effective one-leg operator. Then the process MtM_{t} is expressed in terms of ht,h_{t}, ξt,\xi_{t}, and qk(t)q_{k}(t) as

(8.14) Mt=ht(ξt)λjht(qj(t))λj(ht(ξt)ht(qj(t))ξtqj(t))aβjj<k(ht(qj(t))ht(qk(t))qj(t)qk(t))βjβk.M_{t}=h_{t}^{\prime}(\xi_{t})^{\lambda}\prod_{j}h_{t}^{\prime}(q_{j}(t))^{\lambda_{j}}\Big{(}\frac{h_{t}(\xi_{t})-h_{t}(q_{j}(t))}{\xi_{t}-q_{j}(t)}\Big{)}^{a\beta_{j}}\prod_{j<k}\Big{(}\frac{h_{t}(q_{j}(t))-h_{t}(q_{k}(t))}{q_{j}(t)-q_{k}(t)}\Big{)}^{\beta_{j}\beta_{k}}.

We now present the conformal field theoretic proof for the restriction formula ([13]) of chordal SLE[𝜷]:\mathrm{SLE}[\boldsymbol{\beta}]:

(8.15) dMt=c6Sht(ξt)Mtdt+martingale terms.\mathrm{d}M_{t}=\frac{c}{6}S_{h_{t}}(\xi_{t})M_{t}\,\mathrm{d}t+\textrm{martingale terms.}

Let

F(z,𝒒,t):=G(z,𝒒,t)H(z,𝒒),{G(z,𝒒,t):=𝐄(ΛΩteffid)(z,𝒒)(=𝐄(Λeffht1)(z,𝒒)),H(z,𝒒):=𝐄(Λeffid)(z,𝒒).F(z,\boldsymbol{q},t):=\frac{G(z,\boldsymbol{q},t)}{H(z,\boldsymbol{q})},\quad\begin{cases}G(z,\boldsymbol{q},t):=&\mathbf{E}\,(\Lambda^{\mathrm{eff}}_{\Omega_{t}}\|\mathrm{id})(z,\boldsymbol{q})(=\mathbf{E}\,(\Lambda^{\mathrm{eff}}_{\mathbb{H}}\|h_{t}^{-1})(z,\boldsymbol{q})),\\ H(z,\boldsymbol{q}):=&\mathbf{E}\,(\Lambda^{\mathrm{eff}}_{\mathbb{H}}\|\mathrm{id})(z,\boldsymbol{q}).\end{cases}

Recall that the driving process ξt\xi_{t} satisfies

dξt=κBt+κHξ(ξt,𝒒(t))H(ξt,𝒒(t))dt.\mathrm{d}\xi_{t}=\sqrt{\kappa}B_{t}+\kappa\frac{H_{\xi}(\xi_{t},\boldsymbol{q}(t))}{H(\xi_{t},\boldsymbol{q}(t))}\,\mathrm{d}t.

It follows from Itô’s formula that the drift term of dMt/Mt\mathrm{d}M_{t}/M_{t} equals

(F˙F+κ2FξξF+jFqjFq˙j(t)+κFξFHξH)dt\Big{(}\frac{\dot{F}}{F}+\frac{\kappa}{2}\frac{F_{\xi\xi}}{F}+\sum_{j}\frac{F_{q_{j}}}{F}\dot{q}_{j}(t)+\kappa\frac{F_{\xi}}{F}\frac{H_{\xi}}{H}\Big{)}\,\mathrm{d}t

evaluated at (ξt,𝒒(t),t).(\xi_{t},\boldsymbol{q}(t),t). We rewrite the drift term of dMt/Mt\mathrm{d}M_{t}/M_{t} in terms of GG and H:H:

the drift term of dMtMt=(G˙G+κ2(GξξGHξξH)+j(GqjGHqjH)2qj(t)ξt)dt.\textrm{the drift term of }\frac{\mathrm{d}M_{t}}{M_{t}}=\Big{(}\frac{\dot{G}}{G}+\frac{\kappa}{2}\Big{(}\frac{G_{\xi\xi}}{G}-\frac{H_{\xi\xi}}{H}\Big{)}+\sum_{j}\Big{(}\frac{G_{q_{j}}}{G}-\frac{H_{q_{j}}}{H}\Big{)}\frac{2}{q_{j}(t)-\xi_{t}}\Big{)}\,\mathrm{d}t.

Here we use

F˙F=G˙G,FqjF=GqjGHqjH,FξξF=GξξGHξξH2FξFHξH.\frac{\dot{F}}{F}=\frac{\dot{G}}{G},\qquad\frac{F_{q_{j}}}{F}=\frac{G_{q_{j}}}{G}-\frac{H_{q_{j}}}{H},\qquad\frac{F_{\xi\xi}}{F}=\frac{G_{\xi\xi}}{G}-\frac{H_{\xi\xi}}{H}-2\frac{F_{\xi}}{F}\frac{H_{\xi}}{H}.

By Ward’s equations we have

G˙(z,𝒒(t),t)=\displaystyle\dot{G}(z,\boldsymbol{q}(t),t)= 2ht(ξt)2ht(z)λjht(qj(t))λj𝐄(A(ht(ξt))Λeff(ht(z),ht(q1(t)),)id)\displaystyle-2h_{t}^{\prime}(\xi_{t})^{2}h_{t}^{\prime}(z)^{\lambda}\prod_{j}h_{t}^{\prime}(q_{j}(t))^{\lambda_{j}}\mathbf{E}\big{(}A_{\mathbb{H}}(h_{t}(\xi_{t}))\Lambda^{\mathrm{eff}}_{\mathbb{H}}(h_{t}(z),h_{t}(q_{1}(t)),\cdots)\|\mathrm{id}\big{)}
+2𝐄(vξtΛΩteffid)(z,𝒒(t)).\displaystyle+2\mathbf{E}\,(\mathcal{L}_{v_{\xi_{t}}}\Lambda^{\mathrm{eff}}_{\Omega_{t}}\|\mathrm{id})(z,\boldsymbol{q}(t)).

It follows from conformal invariance that

G˙(z,𝒒(t),t)=2(𝐄AΩt(ξt)ΛΩteff(z,𝒒(t))id)+2𝐄(kξtΛΩteffid)(z,𝒒(t)).\dot{G}(z,\boldsymbol{q}(t),t)=-2(\mathbf{E}\,A_{\Omega_{t}}(\xi_{t})\Lambda^{\mathrm{eff}}_{\Omega_{t}}(z,\boldsymbol{q}(t))\|\mathrm{id})+2\mathbf{E}\,(\mathcal{L}_{k_{\xi_{t}}}\Lambda^{\mathrm{eff}}_{\Omega_{t}}\|\mathrm{id})(z,\boldsymbol{q}(t)).

Sending zz to ξt,\xi_{t},

G˙(ξt,𝒒(t),t)=2(𝐄AΩtξΛΩteff(ξt,𝒒(t))id)+2𝐄(ˇkξtΛΩteffid)(ξt,𝒒(t)),\dot{G}(\xi_{t},\boldsymbol{q}(t),t)=-2(\mathbf{E}\,A_{\Omega_{t}}*_{\xi}\Lambda^{\mathrm{eff}}_{\Omega_{t}}(\xi_{t},\boldsymbol{q}(t))\|\mathrm{id})+2\mathbf{E}\,(\check{\mathcal{L}}_{k_{\xi_{t}}}\Lambda^{\mathrm{eff}}_{\Omega_{t}}\|\mathrm{id})(\xi_{t},\boldsymbol{q}(t)),

where ˇkξt\check{\mathcal{L}}_{k_{\xi_{t}}} does not apply to ξt.\xi_{t}. By the level two degeneracy equation for Λeff\Lambda^{\mathrm{eff}} with respect to ξ\xi,

κ2(GξξGHξξH)=2(𝐄TΩtξΛΩteff(ξt,𝒒(t))id)(𝐄ΛΩteff(ξt,𝒒(t))id)2(𝐄TξΛeff(ξt,𝒒(t))id)(𝐄Λeff(ξt,𝒒(t))id).\frac{\kappa}{2}\Big{(}\frac{G_{\xi\xi}}{G}-\frac{H_{\xi\xi}}{H}\Big{)}=2\frac{(\mathbf{E}\,T_{\Omega_{t}}*_{\xi}\Lambda^{\mathrm{eff}}_{\Omega_{t}}(\xi_{t},\boldsymbol{q}(t))\|\mathrm{id})}{(\mathbf{E}\,\Lambda^{\mathrm{eff}}_{\Omega_{t}}(\xi_{t},\boldsymbol{q}(t))\|\mathrm{id})}-2\frac{(\mathbf{E}\,T_{\mathbb{H}}*_{\xi}\Lambda^{\mathrm{eff}}_{\mathbb{H}}(\xi_{t},\boldsymbol{q}(t))\|\mathrm{id})}{(\mathbf{E}\,\Lambda^{\mathrm{eff}}_{\mathbb{H}}(\xi_{t},\boldsymbol{q}(t))\|\mathrm{id})}.

It follows from Ward’s equation that

(𝐄TξΛeff(ξt,𝒒(t))id)=𝐄(ˇkξtΛeffid)(ξt,𝒒(t)).(\mathbf{E}\,T_{\mathbb{H}}*_{\xi}\Lambda^{\mathrm{eff}}_{\mathbb{H}}(\xi_{t},\boldsymbol{q}(t))\|\mathrm{id})=\mathbf{E}\,(\check{\mathcal{L}}_{k_{\xi_{t}}}\Lambda^{\mathrm{eff}}_{\mathbb{H}}\|\mathrm{id})(\xi_{t},\boldsymbol{q}(t)).

The formula for Lie derivatives of differentials gives

𝐄(ˇkξtΛΩteffid)(ξt,𝒒(t))(𝐄ΛΩteff(ξt,𝒒(t))id)𝐄(ˇkξtΛeffid)(ξt,𝒒(t))(𝐄Λeff(ξt,𝒒(t))id)=j(GqjGHqjH)1qj(t)ξt.\frac{\mathbf{E}\,(\check{\mathcal{L}}_{k_{\xi_{t}}}\Lambda^{\mathrm{eff}}_{\Omega_{t}}\|\mathrm{id})(\xi_{t},\boldsymbol{q}(t))}{(\mathbf{E}\,\Lambda^{\mathrm{eff}}_{\Omega_{t}}(\xi_{t},\boldsymbol{q}(t))\|\mathrm{id})}-\frac{\mathbf{E}\,(\check{\mathcal{L}}_{k_{\xi_{t}}}\Lambda^{\mathrm{eff}}_{\mathbb{H}}\|\mathrm{id})(\xi_{t},\boldsymbol{q}(t))}{(\mathbf{E}\,\Lambda^{\mathrm{eff}}_{\mathbb{H}}(\xi_{t},\boldsymbol{q}(t))\|\mathrm{id})}=-\sum_{j}\Big{(}\frac{G_{q_{j}}}{G}-\frac{H_{q_{j}}}{H}\Big{)}\frac{1}{q_{j}(t)-\xi_{t}}.

Combining all of the above, we have

the drift term of dMtMt\displaystyle\textrm{the drift term of }\frac{\mathrm{d}M_{t}}{M_{t}} =2(𝐄AΩtξΛΩteff(ξt,𝒒(t))id)(𝐄ΛΩteff(ξt,𝒒(t))id)dt+2(𝐄TΩtξΛΩteff(ξt,𝒒(t))id)(𝐄ΛΩteff(ξt,𝒒(t))id)dt\displaystyle=-2\frac{(\mathbf{E}\,A_{\Omega_{t}}*_{\xi}\Lambda^{\mathrm{eff}}_{\Omega_{t}}(\xi_{t},\boldsymbol{q}(t))\|\mathrm{id})}{(\mathbf{E}\,\Lambda^{\mathrm{eff}}_{\Omega_{t}}(\xi_{t},\boldsymbol{q}(t))\|\mathrm{id})}\,\mathrm{d}t+2\frac{(\mathbf{E}\,T_{\Omega_{t}}*_{\xi}\Lambda^{\mathrm{eff}}_{\Omega_{t}}(\xi_{t},\boldsymbol{q}(t))\|\mathrm{id})}{(\mathbf{E}\,\Lambda^{\mathrm{eff}}_{\Omega_{t}}(\xi_{t},\boldsymbol{q}(t))\|\mathrm{id})}\,\mathrm{d}t
=c6Sht(ξt)dt.\displaystyle=\frac{c}{6}S_{h_{t}}(\xi_{t})\,\mathrm{d}t.

We now present a conformal field theoretic proof for the restriction formula of radial SLE[𝜷]\mathrm{SLE}[\boldsymbol{\beta}] (κ=8/3).(\kappa=8/3). For this purpose, we let ζ,qkD,qD,\zeta,q_{k}\in\partial D,q\in D,

𝜷=𝝉+𝜷ˇ,𝝉=aζa2qa2q𝜷ˇ=βkqk+(bβ2)q+(bβ2)q,β=βk,\boldsymbol{\beta}=\boldsymbol{\tau}+\check{\boldsymbol{\beta}\,},\quad\boldsymbol{\tau}=a\cdot\zeta-\frac{a}{2}\cdot q-\frac{a}{2}\cdot q^{*}\quad\check{\boldsymbol{\beta}\,}=\sum\beta_{k}\cdot q_{k}+(b-\frac{\beta}{2})\cdot q+(b-\frac{\beta}{2})\cdot q^{*},\quad\beta=\sum\beta_{k},

and consider the effective one-leg operator Λeff,\Lambda^{\mathrm{eff}},

Λeff(ζ):=𝒫𝜷ˇΛ(ζ)=C(b)[𝜷]eiΦ[𝝉].\Lambda^{\mathrm{eff}}(\zeta):=\mathcal{P}_{\check{\boldsymbol{\beta}\,}}\,\Lambda(\zeta)=C_{(b)}[\boldsymbol{\beta}]\,\mathrm{e}^{\odot i\Phi[\boldsymbol{\tau}]}.

Fix a hull KK such that KK does not intersect ζ=1,q1,,qk,.\zeta=1,q_{1},\cdots,q_{k},\cdots. Let gtg_{t} be the radial SLE[𝜷]\mathrm{SLE}[\boldsymbol{\beta}] map with the hull Kt.K_{t}. On the event KK=,K_{\infty}\cap K=\emptyset, a conformal map ht:Ωt=gt(DtK)𝔻h_{t}:\Omega_{t}=g_{t}(D_{t}\setminus K)\to\mathbb{D} is defined by

ht=g~tΨKgt1,h_{t}=\widetilde{g}_{t}\circ\Psi_{K}\circ g_{t}^{-1},

where g~t\widetilde{g}_{t} is a radial Loewner map from D~t=DK~t\widetilde{D}_{t}=D\setminus\widetilde{K}_{t} onto 𝔻,\mathbb{D}, K~t=ΨK(Kt),\widetilde{K}_{t}=\Psi_{K}(K_{t}), and ΨK\Psi_{K} is the conformal map (𝔻K,0)(𝔻,0)(\mathbb{D}\setminus K,0)\to(\mathbb{D},0) satisfying ΨK(0)>0.\Psi_{K}^{\prime}(0)>0. Let

(8.16) Mt:=(Z𝜷tidΩt)(Z𝜷tid),𝜷t=aζt+(ba+β2)q+(ba+β2)q+βkqk(t),M_{t}:=\frac{(Z_{\boldsymbol{\beta}_{t}}\|\,\mathrm{id}_{\Omega_{t}})}{(Z_{\boldsymbol{\beta}_{t}}\|\,\mathrm{id}_{\mathbb{H}})},\quad\boldsymbol{\beta}_{t}=a\cdot\zeta_{t}+(b-\frac{a+\beta}{2})\cdot q+(b-\frac{a+\beta}{2})\cdot q^{*}+\sum\beta_{k}\cdot q_{k}(t),

where qk(t)q_{k}(t) satisfies qk(0)=qkq_{k}(0)=q_{k} and

dqk(t)=qk(t)ζt+qk(t)ζtqk(t)dt.\mathrm{d}q_{k}(t)=q_{k}(t)\frac{\zeta_{t}+q_{k}(t)}{\zeta_{t}-q_{k}(t)}\,\mathrm{d}t.

We now represent MtM_{t} in terms of the effective one-leg operator:

Mt=F(ζt,𝒒(t),t),F(z,𝒒,t)=G(z,𝒒,t)H(z,𝒒),M_{t}=F(\zeta_{t},\boldsymbol{q}(t),t),\quad F(z,\boldsymbol{q},t)=\frac{G(z,\boldsymbol{q},t)}{H(z,\boldsymbol{q})},

where G(z,𝒒,t)=𝐄(Λ𝔻effht1)(z),G(z,\boldsymbol{q},t)=\mathbf{E}(\Lambda_{\mathbb{D}}^{\mathrm{eff}}\|h_{t}^{-1})(z), and H(z,𝒒)=𝐄(Λ𝔻effid)(z).H(z,\boldsymbol{q})=\mathbf{E}(\Lambda_{\mathbb{D}}^{\mathrm{eff}}\|\mathrm{id})(z).

Theorem 8.4.

We have

(8.17) the drift term of dMtMt=c6ζt2Sht(ζt)dt.\textrm{the drift term of }\frac{\mathrm{d}M_{t}}{M_{t}}=-\frac{c}{6}\,\zeta_{t}^{2}\,S_{h_{t}}(\zeta_{t})\,\mathrm{d}t.
Proof.

Recall that

dζt=iκζtdBtκ2ζtdtκζt2(HζH+hζt)dt=iκζtdBtκζt2HζHdt3ζtdt.\mathrm{d}\zeta_{t}=i\sqrt{\kappa}\,\zeta_{t}\,\mathrm{d}B_{t}-\frac{\kappa}{2}\zeta_{t}\,\mathrm{d}t-\kappa\zeta_{t}^{2}\Big{(}\frac{H_{\zeta}}{H}+\frac{h}{\zeta_{t}}\Big{)}\,\mathrm{d}t=i\sqrt{\kappa}\,\zeta_{t}\,\mathrm{d}B_{t}-\kappa\zeta_{t}^{2}\frac{H_{\zeta}}{H}\,\mathrm{d}t-3\zeta_{t}\,\mathrm{d}t.

It follows from Itô’s formula that

dMt=F˙dtκ2ζt2Fζζdt+Fζdζt+Fqkdqk(t)\mathrm{d}M_{t}=\dot{F}\,\mathrm{d}t-\frac{\kappa}{2}\zeta_{t}^{2}\,F_{\zeta\zeta}\,\mathrm{d}t+F_{\zeta}\,\mathrm{d}\zeta_{t}+\sum F_{q_{k}}\,\mathrm{d}q_{k}(t)

evaluated at (ζt,𝒒(t),t).(\zeta_{t},\boldsymbol{q}(t),t). The drift term of dMt/Mt\mathrm{d}M_{t}/M_{t} is

(F˙Fκ2ζt2FζζF(κζt2HζH+3ζt)FζF+k(qk(t)ζt+qk(t)ζtqk(t))FqkF)|(ζt,𝒒(t),t).\Big{(}\frac{\dot{F}}{F}-\frac{\kappa}{2}\zeta_{t}^{2}\,\frac{F_{\zeta\zeta}}{F}-\Big{(}\kappa\zeta_{t}^{2}\frac{H_{\zeta}}{H}+3\zeta_{t}\Big{)}\frac{F_{\zeta}}{F}+\sum_{k}\Big{(}q_{k}(t)\frac{\zeta_{t}+q_{k}(t)}{\zeta_{t}-q_{k}(t)}\Big{)}\frac{F_{q_{k}}}{F}\Big{)}\Big{|}_{(\zeta_{t},\boldsymbol{q}(t),t)}.

We rewrite the drift term of dMt/Mt\mathrm{d}M_{t}/M_{t} in terms of GG and H:H:

(8.18) (G˙Gκ2ζt2(GζζGHζζH)3ζt(GζGHζH)+k(qk(t)ζt+qk(t)ζtqk(t))(GqkGHqkH))\Big{(}\frac{\dot{G}}{G}-\frac{\kappa}{2}\zeta_{t}^{2}\,\Big{(}\frac{G_{\zeta\zeta}}{G}-\frac{H_{\zeta\zeta}}{H}\Big{)}-3\zeta_{t}\Big{(}\frac{G_{\zeta}}{G}-\frac{H_{\zeta}}{H}\Big{)}+\sum_{k}\Big{(}q_{k}(t)\frac{\zeta_{t}+q_{k}(t)}{\zeta_{t}-q_{k}(t)}\Big{)}\Big{(}\frac{G_{q_{k}}}{G}-\frac{H_{q_{k}}}{H}\Big{)}\Big{)}

evaluated at (ξt,𝒒(t),t).(\xi_{t},\boldsymbol{q}(t),t).

Using the similar method in [18, Section 14.5], we represent G˙\dot{G} in terms of the Lie derivatives:

(8.19) G˙(z,𝒒,t)\displaystyle\dot{G}(z,\boldsymbol{q},t) =dds|s=0(𝐄Λ𝔻effht+s1)(z)=dds|s=0(𝐄Λ𝔻effht1fs,t1)(z)\displaystyle=\frac{\mathrm{d}}{\mathrm{d}s}\Big{|}_{s=0}(\mathbf{E}\,\Lambda^{\mathrm{eff}}_{\mathbb{D}}\,\|\,h_{t+s}^{-1})(z)=\frac{\mathrm{d}}{\mathrm{d}s}\Big{|}_{s=0}(\mathbf{E}\,\Lambda^{\mathrm{eff}}_{\mathbb{D}}\,\|\,h_{t}^{-1}\circ f_{s,t}^{-1})(z)
=(𝐄(v,𝔻)Λ𝔻effht1)(z),\displaystyle=(\mathbf{E}\,\mathcal{L}(v,\mathbb{D})\,\Lambda^{\mathrm{eff}}_{\mathbb{D}}\,\|\,h_{t}^{-1})(z),

where fs,t=ht+sht1f_{s,t}=h_{t+s}\circ h_{t}^{-1} and

(vid𝔻)=dds|s=0fs,t=h˙tht1.(v\,\|\,\mathrm{id}_{\mathbb{D}})=\frac{\mathrm{d}}{\mathrm{d}s}\Big{|}_{s=0}f_{s,t}=\dot{h}_{t}\circ h_{t}^{-1}.

We only need to compute the vector field v.v. We represent vv as the difference of two Loewner vector fields associated with the flows in the domains 𝔻\mathbb{D} and Ωt.\Omega_{t}. Applying the chain rule to ht=g~tΨKgt1h_{t}=\widetilde{g}_{t}\circ\Psi_{K}\circ g_{t}^{-1} and computing the capacity changes, we have

h˙t(z)=|ht(ζt)|2vζ~t(ht(z))ht(z)vζt(z),(vζ(z)=zζ+zζz),\dot{h}_{t}(z)=|h_{t}^{\prime}(\zeta_{t})|^{2}v_{\widetilde{\zeta}_{t}}(h_{t}(z))-h_{t}^{\prime}(z)v_{\zeta_{t}}(z),\quad\Big{(}v_{\zeta}(z)=z\frac{\zeta+z}{\zeta-z}\Big{)},

where ζ~t=ht(ζt).\widetilde{\zeta}_{t}=h_{t}(\zeta_{t}). By the above equation and (vid𝔻)=h˙tht1,(v\,\|\,\mathrm{id}_{\mathbb{D}})=\dot{h}_{t}\circ h_{t}^{-1},

(8.20) (vid𝔻)(z)=|ht(ζt)|2vζ~t(z)ht(ht1(z))vζt(ht1(z)).(v\,\|\,\mathrm{id}_{\mathbb{D}})(z)=|h_{t}^{\prime}(\zeta_{t})|^{2}v_{\widetilde{\zeta}_{t}}(z)-h_{t}^{\prime}(h_{t}^{-1}(z))v_{\zeta_{t}}(h_{t}^{-1}(z)).

It follows from (8.19) and (8.20) that

G˙(z,𝒒(t),t)\displaystyle\dot{G}(z,\boldsymbol{q}(t),t) =|ht(ζt)|2ht(z)λ(ht(qk(t)))λk𝐄((vζ~t,𝔻)Λ𝔻eff(ht(z))id𝔻)\displaystyle=|h_{t}^{\prime}(\zeta_{t})|^{2}h_{t}^{\prime}(z)^{\lambda}\prod\big{(}h_{t}^{\prime}(q_{k}(t))\big{)}^{\lambda_{k}}\,\mathbf{E}\,\big{(}\mathcal{L}(v_{\widetilde{\zeta}_{t}},\mathbb{D})\,\Lambda^{\mathrm{eff}}_{\mathbb{D}}(h_{t}(z)\big{)}\,\|\,\mathrm{id}_{\mathbb{D}})
(𝐄(vζt)ΛΩteffidΩt)(z).\displaystyle-\big{(}\mathbf{E}\,\mathcal{L}(v_{\zeta_{t}})\,\Lambda^{\mathrm{eff}}_{\Omega_{t}}\,\|\,\mathrm{id}_{\Omega_{t}}\big{)}(z).

By Ward’s equation,

G˙(z,𝒒(t),t)\displaystyle\dot{G}(z,\boldsymbol{q}(t),t) =2|ht(ζt)|2(ht(qk(t)))λkht(z)λζ~t2(𝐄T𝔻(ζ~t)Λ𝔻eff(ht(z))id𝔻)\displaystyle=2|h_{t}^{\prime}(\zeta_{t})|^{2}\prod\big{(}h_{t}^{\prime}(q_{k}(t))\big{)}^{\lambda_{k}}h_{t}^{\prime}(z)^{\lambda}\widetilde{\zeta}_{t}^{2}(\mathbf{E}\,T_{\mathbb{D}}(\widetilde{\zeta}_{t})\,\Lambda^{\mathrm{eff}}_{\mathbb{D}}(h_{t}(z))\,\|\,\mathrm{id}_{\mathbb{D}})
(𝐄(vζt)ΛΩteffidΩt)(z).\displaystyle-(\mathbf{E}\,\mathcal{L}(v_{\zeta_{t}})\,\Lambda^{\mathrm{eff}}_{\Omega_{t}}\,\|\,\mathrm{id}_{\Omega_{t}}\big{)}(z).

It follows from conformal invariance that

G˙(z,𝒒(t),t)\displaystyle\dot{G}(z,\boldsymbol{q}(t),t) =2ζt2(𝐄TΩt(ζt)ΛΩteff(z)idΩt)c6ζt2Sht(ζt)(𝐄ΛΩteffidΩt)(z)\displaystyle=2\zeta_{t}^{2}(\mathbf{E}\,T_{\Omega_{t}}(\zeta_{t})\,\Lambda^{\mathrm{eff}}_{\Omega_{t}}(z)\,\|\,\mathrm{id}_{\Omega_{t}})-\frac{c}{6}\zeta_{t}^{2}S_{h_{t}}(\zeta_{t})\,(\mathbf{E}\,\Lambda^{\mathrm{eff}}_{\Omega_{t}}\,\|\,\mathrm{id}_{\Omega_{t}}\big{)}(z)
(𝐄(vζt)ΛΩteffidΩt)(z).\displaystyle-(\mathbf{E}\,\mathcal{L}(v_{\zeta_{t}})\,\Lambda^{\mathrm{eff}}_{\Omega_{t}}\,\|\,\mathrm{id}_{\Omega_{t}}\big{)}(z).

Sending zz to ζt\zeta_{t} and applying (7.15) (and T1Λeff=Λeff,T2Λeff=hΛeffT*_{-1}\Lambda^{\mathrm{eff}}=\partial\Lambda^{\mathrm{eff}},T*_{-2}\Lambda^{\mathrm{eff}}=h\Lambda^{\mathrm{eff}}, see (7.16)),

G˙(ζt,𝒒(t),t)G(ζt,𝒒(t),t)=c6ζt2Sht(ζt)+L(ζt)G(ζt,𝒒(t),t)G(ζt,𝒒(t),t)ˇvζtG(ζt,𝒒(t),t)G(ζt,𝒒(t),t),\frac{\dot{G}(\zeta_{t},\boldsymbol{q}(t),t)}{G(\zeta_{t},\boldsymbol{q}(t),t)}=-\frac{c}{6}\zeta_{t}^{2}S_{h_{t}}(\zeta_{t})+\frac{L(\zeta_{t})G(\zeta_{t},\boldsymbol{q}(t),t)}{G(\zeta_{t},\boldsymbol{q}(t),t)}-\frac{\check{\mathcal{L}}_{v_{\zeta_{t}}}G(\zeta_{t},\boldsymbol{q}(t),t)}{G(\zeta_{t},\boldsymbol{q}(t),t)},

where L(z)=2z2L2(z)+3zL1(z)+L0(z).L(z)=2z^{2}L_{-2}(z)+3zL_{-1}(z)+L_{0}(z). By (8.18) and the level two degeneracy equation, the drift of dMt/Mt\mathrm{d}M_{t}/M_{t} simplifies to

c6ζt2Sht(ζt)\displaystyle-\frac{c}{6}\zeta_{t}^{2}S_{h_{t}}(\zeta_{t}) ˇvζtG(ζt,𝒒(t),t)G(ζt,𝒒(t),t)+(κ2ζt2HζζH+3ζtHH+h)|(ζt,𝒒(t),t)\displaystyle-\frac{\check{\mathcal{L}}_{v_{\zeta_{t}}}G(\zeta_{t},\boldsymbol{q}(t),t)}{G(\zeta_{t},\boldsymbol{q}(t),t)}+\Big{(}\frac{\kappa}{2}\zeta_{t}^{2}\,\frac{H_{\zeta\zeta}}{H}+3\zeta_{t}\frac{H^{\prime}}{H}+h\Big{)}\Big{|}_{(\zeta_{t},\boldsymbol{q}(t),t)}
+k(qk(t)ζt+qk(t)ζtqk(t))(GqkGHqkH)|(ζt,𝒒(t),t).\displaystyle+\sum_{k}\Big{(}q_{k}(t)\frac{\zeta_{t}+q_{k}(t)}{\zeta_{t}-q_{k}(t)}\Big{)}\Big{(}\frac{G_{q_{k}}}{G}-\frac{H_{q_{k}}}{H}\Big{)}\Big{|}_{(\zeta_{t},\boldsymbol{q}(t),t)}.

It follows from the null vector equation that

κ2ζt2HζζH+3ζtHH+h=ˇvζtHH.\frac{\kappa}{2}\zeta_{t}^{2}\,\frac{H_{\zeta\zeta}}{H}+3\zeta_{t}\frac{H^{\prime}}{H}+h=\frac{\check{\mathcal{L}}_{v_{\zeta_{t}}}H}{H}.

Thus we find the drift term of dMt/Mt\mathrm{d}M_{t}/M_{t} as

ζt2c6Sht(ζt)(ˇvζtGGˇvζtHH)+k(qk(t)ζt+qk(t)ζtqk(t))(GqkGHqkH)-\zeta_{t}^{2}\frac{c}{6}S_{h_{t}}(\zeta_{t})-\Big{(}\frac{\check{\mathcal{L}}_{v_{\zeta_{t}}}G}{G}-\frac{\check{\mathcal{L}}_{v_{\zeta_{t}}}H}{H}\Big{)}+\sum_{k}\Big{(}q_{k}(t)\frac{\zeta_{t}+q_{k}(t)}{\zeta_{t}-q_{k}(t)}\Big{)}\Big{(}\frac{G_{q_{k}}}{G}-\frac{H_{q_{k}}}{H}\Big{)}

evaluated at (ξt,𝒒(t),t).(\xi_{t},\boldsymbol{q}(t),t). Since

ˇvζtGGˇvζtHH=k(qk(t)ζt+qk(t)ζtqk(t))(GqkGHqkH),\frac{\check{\mathcal{L}}_{v_{\zeta_{t}}}G}{G}-\frac{\check{\mathcal{L}}_{v_{\zeta_{t}}}H}{H}=\sum_{k}\Big{(}q_{k}(t)\frac{\zeta_{t}+q_{k}(t)}{\zeta_{t}-q_{k}(t)}\Big{)}\Big{(}\frac{G_{q_{k}}}{G}-\frac{H_{q_{k}}}{H}\Big{)},

we have (8.17). ∎

9. Conformal field theory with Neumann boundary condition

In this section we briefly implement a version of conformal field theory from background charge modifications of the Gaussian field with Neumann boundary condition. In the last subsection we present the connection of this theory to the theory of the backward SLEs.

9.1. Gaussian free field with Neumann boundary condition

The Gaussian free field NN in a planar domain DD with Neumann boundary condition is an isometry N:N(D)L2(Ω,𝐏)N:\mathcal{E}_{N}(D)\to L^{2}(\Omega,\mathbf{P}) from the Neumann energy space N(D)\mathcal{E}_{N}(D) such that the image consists of centered Gaussian random variables. Here (Ω,𝐏)(\Omega,\mathbf{P}) is a probability space and N(D)\mathcal{E}_{N}(D) is the completion of smooth functions up to the boundary with mean zero (or the neutrality condition (NC0)(\mathrm{NC}_{0})), compact supports in DD,D\,\cup\,\partial D, and Neumann boundary condition fn=0\nabla f\cdot n=0 (where nn is normal to D\partial D) with respect to the norm

fN2=2GN(ζ,z)f(ζ)f(z)¯dA(ζ)dA(z),\|f\|^{2}_{\mathcal{E}_{N}}=\iint 2G_{N}(\zeta,z)\,f(\zeta)\,\overline{f(z)}~{}\mathrm{d}A(\zeta)\,\mathrm{d}A(z),

where AA is the normalized area measure and GNG_{N} is the Neumann Green’s function for D.D. In the upper half-plane, we have

GN(ζ,z)=log1|(ζz)(ζz¯)|.G_{N}(\zeta,z)=\log\frac{1}{|(\zeta-z)(\zeta-\bar{z})|}.

As in the Dirichlet case, NN can be constructed from the Gaussian free field Ψ\Psi on its Schottky double S=Ddouble.S=D^{\mathrm{double}}. The Neumann energy space N(D)\mathcal{E}_{N}(D) can be embedded isometrically into (S)\mathcal{E}(S) in a natural way. For example, for μ\mu in the energy space N(),\mathcal{E}_{N}(\mathbb{H}),

μN()2\displaystyle\|\mu\|_{\mathcal{E}_{N}(\mathbb{H})}^{2} =×log1|zw|2μ(z)μ(w)¯+×log1|zw¯|2μ(z)μ(w)¯\displaystyle=\int_{\mathbb{H}\times\mathbb{H}}\log\frac{1}{|z-w|^{2}}\,\mu(z)\overline{\mu(w)}+\int_{\mathbb{H}\times\mathbb{H}}\log\frac{1}{|z-\bar{w}|^{2}}\,\mu(z)\overline{\mu(w)}
=12×log1|zw|2ν(z)ν(w)¯=12ν(^)2,\displaystyle=\frac{1}{2}\int_{\mathbb{C}\times\mathbb{C}}\log\frac{1}{|z-w|^{2}}\,\nu(z)\overline{\nu(w)}=\frac{1}{2}\|\nu\|_{\mathcal{E}(\widehat{\mathbb{C}})}^{2},

where ν=μ\nu=\mu in the upper half-plane \mathbb{H} and ν=μ\nu=\mu^{*} in the lower half-plane \mathbb{H}^{*} (μ(E)=μ(E),\mu^{*}(E)=\mu(E^{*}), E={z¯,zE}E^{*}=\{\bar{z},z\in E\}). For example, a test function ff in \mathbb{H} with mean zero and Neumann boundary condition extends smoothly to \mathbb{C} such that f(z¯)=f(z).f(\bar{z})=f(z).

As a Fock space space, the Gaussian free field NNDN\equiv N_{D} in DD can be constructed from the Gaussian free field ΨΨS:\Psi\equiv\Psi_{S}:

N(z)=12(Ψ(z)+Ψ(z)).N(z)=\frac{1}{\sqrt{2}}(\Psi(z)+\Psi(z^{*})).

This field N(z)N(z) is formal and thus we need to require the neutrality condition (NC0)(\mathrm{NC}_{0}) on N.N. For example, N(z,z0):=N(z)N(z0)N(z,z_{0}):=N(z)-N(z_{0}) is a 2-variant well-defined Fock space field. The formal field NN has the formal correlation:

𝐄N(z)N(z0)=2GN(z,z0).\mathbf{E}\,N(z)N(z_{0})=2G_{N}(z,z_{0}).

From the Schottky double construction, the current field JDD=ΦDJ_{D}^{D}=\partial\Phi_{D} in DD with Dirichlet boundary condition and the current field JS=ΨSJ_{S}=\partial\Psi_{S} are related as follows. For example,

JD(z)=12(J^(z)J^(z¯)¯).J_{\mathbb{H}}^{D}(z)=\frac{1}{\sqrt{2}}\big{(}J_{\widehat{\mathbb{C}}}(z)-\overline{J_{\widehat{\mathbb{C}}}(\bar{z})}\big{)}.

On the other hand, the current field JNJ_{\mathbb{H}}^{N} with Neumann boundary condition is related to JJ_{\mathbb{C}} as

JN(z)=12(J^(z)+J^(z¯)¯).J_{\mathbb{H}}^{N}(z)=\frac{1}{\sqrt{2}}\big{(}J_{\widehat{\mathbb{C}}}(z)+\overline{J_{\widehat{\mathbb{C}}}(\bar{z})}\big{)}.

In a similar way, stress tensors are related as

AD(z)=12JDJD(z)=12(A^(z)+A^(z¯)¯+J^(z)J^(z¯)¯)A_{\mathbb{H}}^{D}(z)=-\frac{1}{2}J_{\mathbb{H}}^{D}\odot J_{\mathbb{H}}^{D}(z)=\frac{1}{2}\big{(}A_{\widehat{\mathbb{C}}}(z)+\overline{A_{\widehat{\mathbb{C}}}(\bar{z})}+J_{\widehat{\mathbb{C}}}(z)\odot\overline{J_{\widehat{\mathbb{C}}}(\bar{z})}\big{)}

and

AN(z)=12JNJN(z)=12(A^(z)+A^(z¯)¯J^(z)J^(z¯)¯).A_{\mathbb{H}}^{N}(z)=-\frac{1}{2}J_{\mathbb{H}}^{N}\odot J_{\mathbb{H}}^{N}(z)=\frac{1}{2}\big{(}A_{\widehat{\mathbb{C}}}(z)+\overline{A_{\widehat{\mathbb{C}}}(\bar{z})}-J_{\widehat{\mathbb{C}}}(z)\odot\overline{J_{\widehat{\mathbb{C}}}(\bar{z})}\big{)}.

It is easy to check that ANA^{N} is a stress tensor for N.N. Indeed, as ζz,\zeta\to z, we have the following Ward’s OPE for N:N:

AN(ζ)N(z)=12JN(ζ)JN(ζ)N(z)=𝐄[JN(ζ)N(z)]JN(ζ)JN(ζ)ζzN(z)ζzA^{N}(\zeta)N(z)=-\frac{1}{2}J^{N}(\zeta)\odot J^{N}(\zeta)N(z)=-\mathbf{E}[J^{N}(\zeta)N(z)]J^{N}(\zeta)\sim\frac{J^{N}(\zeta)}{\zeta-z}\sim\frac{\partial N(z)}{\zeta-z}

in the identity chart of .\mathbb{H}. Here we use

𝐄JN(ζ)N(z)=1ζz1ζz¯\mathbf{E}\,J^{N}(\zeta)N(z)=-\frac{1}{\zeta-z}-\frac{1}{\zeta-\bar{z}}

in .\mathbb{H}.

Formal fields N±(z)N^{\pm}(z) with Neumann boundary condition are defined by

N±(z)=12(Ψ±(z)+Ψ(z)).N^{\pm}(z)=\frac{1}{\sqrt{2}}(\Psi^{\pm}(z)+\Psi^{\mp}(z^{*})).

For example, we have

(9.1) 𝐄N(z)N+(z0)=log1zz0+log1z¯z0,𝐄N(z)N(z0)=log1zz¯0+log1z¯z¯0,\mathbf{E}\,N(z)N^{+}(z_{0})=\log\frac{1}{z-z_{0}}+\log\frac{1}{\bar{z}-z_{0}},\qquad\mathbf{E}\,N(z)N^{-}(z_{0})=\log\frac{1}{z-\bar{z}_{0}}+\log\frac{1}{\bar{z}-\bar{z}_{0}},

and

𝐄N+(z)N+(z0)=log1zz0,𝐄N+(z)N(z0)=log1zz¯0\mathbf{E}\,N^{+}(z)N^{+}(z_{0})=\log\frac{1}{z-z_{0}},\qquad\mathbf{E}\,N^{+}(z)N^{-}(z_{0})=\log\frac{1}{z-\bar{z}_{0}}

in .\mathbb{H}. As in the Dirichlet case, 2-variant fields N±(z,z0):=N±(z)N±(z0)N^{\pm}(z,z_{0}):=N^{\pm}(z)-N^{\pm}(z_{0}) are well-defined multivalued Fock space fields.

9.2. Modifications and OPE exponentials

We first define background charge modifications of the chiral bosonic fields with Neumann boundary condition. Given double background charges 𝜷±=kβk±qk\boldsymbol{\beta}^{\pm}=\sum_{k}\beta_{k}^{\pm}\cdot q_{k} with the neutrality conditions (NCb±)(\mathrm{NC}_{b^{\pm}}) in D,D, we define the formal fields N𝜷±N𝜷+,𝜷±N_{\boldsymbol{\beta}}^{\pm}\equiv N_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}^{\pm} (𝜷=𝜷++𝜷\boldsymbol{\beta}=\boldsymbol{\beta}^{+}+\boldsymbol{\beta}_{*}^{-}) by

N𝜷±:=N±+n𝜷±,N_{\boldsymbol{\beta}}^{\pm}:=N^{\pm}+n_{\boldsymbol{\beta}}^{\pm},

where n𝜷+n𝜷+,𝜷+n_{\boldsymbol{\beta}}^{+}\equiv n_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}^{+} is a PPS(b,0)\mathrm{PPS}(b,0) form (b=b++bb=b^{+}+b^{-}) and n𝜷n𝜷+,𝜷n_{\boldsymbol{\beta}}^{-}\equiv n_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}^{-} is a PPS(0,b)\mathrm{PPS}(0,b) form such that

(9.2) n𝜷+(z)\displaystyle n_{\boldsymbol{\beta}}^{+}(z) =βk+log1zqk+βklog1zq¯k\displaystyle=\sum\beta_{k}^{+}\log\frac{1}{z-q_{k}}+\sum\beta_{k}^{-}\log\frac{1}{z-\bar{q}_{k}}
n𝜷(z)\displaystyle n_{\boldsymbol{\beta}}^{-}(z) =βk+log1z¯qk+βklog1z¯q¯k\displaystyle=\sum\beta_{k}^{+}\log\frac{1}{\bar{z}-q_{k}}+\sum\beta_{k}^{-}\log\frac{1}{\bar{z}-\bar{q}_{k}}

in the \mathbb{H}-uniformization. Furthermore, n𝜷:=n𝜷++n𝜷n_{\boldsymbol{\beta}}:=n_{\boldsymbol{\beta}}^{+}+n_{\boldsymbol{\beta}}^{-} is a PPS(b,b)\mathrm{PPS}(b,b) form. For two divisors 𝝉±\boldsymbol{\tau}^{\pm} satisfying the neutrality condition (NC0),(\mathrm{NC}_{0}), we define

N𝜷[𝝉]N𝜷[𝝉+,𝝉]:=j(τj+N𝜷+(zj)+τjN𝜷(zj)),N_{\boldsymbol{\beta}}[\boldsymbol{\tau}]\equiv N_{\boldsymbol{\beta}}[\boldsymbol{\tau}^{+},\boldsymbol{\tau}^{-}]:=\sum_{j}\big{(}\tau_{j}^{+}N_{\boldsymbol{\beta}}^{+}(z_{j})+\tau_{j}^{-}N_{\boldsymbol{\beta}}^{-}(z_{j})\big{)},

where 𝝉=𝝉++𝝉.\boldsymbol{\tau}=\boldsymbol{\tau}^{+}+\boldsymbol{\tau}_{*}^{-}. Compare this definition to the definition of Φ𝜷[𝝉]\Phi_{\boldsymbol{\beta}}[\boldsymbol{\tau}] in the Dirichlet case:

Φ𝜷[𝝉]:=j(τj+Φ𝜷+(zj)τjΦ𝜷(zj)),\Phi_{\boldsymbol{\beta}}[\boldsymbol{\tau}]:=\sum_{j}\big{(}\tau_{j}^{+}\Phi_{\boldsymbol{\beta}}^{+}(z_{j})-\tau_{j}^{-}\Phi_{\boldsymbol{\beta}}^{-}(z_{j})\big{)},

where 𝝉=𝝉++𝝉\boldsymbol{\tau}=\boldsymbol{\tau}^{+}+\boldsymbol{\tau}_{*}^{-} satisfies the neutrality condition (NC0).(\mathrm{NC}_{0}). We now explain why the signs of τj\tau_{j}^{-} appear differently. The choices of signs are consistent with the representation of 𝝉=𝝉++𝝉\boldsymbol{\tau}=\boldsymbol{\tau}^{+}+\boldsymbol{\tau}_{*}^{-} when the nodes are on the boundary. For zD,z\in\partial D, we have Φ(z)=0\Phi(z)=0 and thus the formal field Φ[1z]\Phi[1\cdot z] can be represented either by Φ+(z)\Phi^{+}(z) or by Φ(z).-\Phi^{-}(z). On the other hand, the formal field N[1z]N[1\cdot z] is represented by N+(z)=N(z).N^{+}(z)=N^{-}(z). The relation N+(z)=N(z)N^{+}(z)=N^{-}(z) for zDz\in\partial D is obvious:

𝐄N+(z)N(z0)=log1(zz0)(zz¯0)=log1(z¯z0)(z¯z¯0)=𝐄N(z)N(z0)\mathbf{E}\,N^{+}(z)N(z_{0})=\log\frac{1}{(z-z_{0})(z-\bar{z}_{0})}=\log\frac{1}{(\bar{z}-z_{0})(\bar{z}-\bar{z}_{0})}=\mathbf{E}\,N^{-}(z)N(z_{0})

in the \mathbb{H}-uniformization.

A stress tensor A𝜷A𝜷NA_{\boldsymbol{\beta}}\equiv A_{\boldsymbol{\beta}}^{N} for N𝜷N_{\boldsymbol{\beta}} is given by

A𝜷=A+(bj𝜷)J,j𝜷:=𝐄J𝜷,J𝜷J𝜷N=N𝜷.A_{\boldsymbol{\beta}}=A+(b\partial-j_{\boldsymbol{\beta}})J,\qquad j_{\boldsymbol{\beta}}:=\mathbf{E}\,J_{\boldsymbol{\beta}},\qquad J_{\boldsymbol{\beta}}\equiv J_{\boldsymbol{\beta}}^{N}=\partial N_{\boldsymbol{\beta}}.

As ζz,\zeta\to z, in ,\mathbb{H}, Ward’s OPE for N𝜷N_{\boldsymbol{\beta}} holds:

A𝜷(ζ)N𝜷(z,z0)=(A(ζ)+(bζj𝜷(ζ))J(ζ))N(z,z0)J𝜷(z)ζz+b(ζz)2.A_{\boldsymbol{\beta}}(\zeta)N_{\boldsymbol{\beta}}(z,z_{0})=\Big{(}A(\zeta)+\big{(}b\partial_{\zeta}-j_{\boldsymbol{\beta}}(\zeta)\big{)}J(\zeta)\Big{)}N(z,z_{0})\\ \sim\frac{J_{\boldsymbol{\beta}}(z)}{\zeta-z}+\frac{b}{(\zeta-z)^{2}}.

In the symmetric case 𝜷=𝜷¯,\boldsymbol{\beta}=\mkern 1.5mu\overline{\mkern-1.5mu\boldsymbol{\beta}_{*}\mkern-1.5mu}\mkern 1.5mu, the Virasoro field

T𝜷=12J𝜷J𝜷+bJ𝜷T_{\boldsymbol{\beta}}=-\frac{1}{2}J_{\boldsymbol{\beta}}*J_{\boldsymbol{\beta}}+b\partial J_{\boldsymbol{\beta}}

gives rise to the Virasoro pair (T𝜷,T𝜷¯).(T_{\boldsymbol{\beta}},\mkern 1.5mu\overline{\mkern-1.5muT_{\boldsymbol{\beta}}\mkern-1.5mu}\mkern 1.5mu). The central charge cc in this theory is given by

c=1+12b2.c=1+12b^{2}.

If the parameter bb is related to the SLE parameter κ\kappa as

(9.3) b=a(κ/4+1),a=±2/κ,b=-a(\kappa/4+1),\qquad a=\pm\sqrt{2/\kappa},

then

c=13+12(κ/8+2/κ)25.c=13+12(\kappa/8+2/\kappa)\geq 25.

Modified multi-vertex fields or OPE exponentials 𝒪𝜷[𝝉]\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}] are defined by

𝒪𝜷[𝝉]=C(ib)[i(𝝉+𝜷)]C(ib)[i𝜷]eN[𝝉].\mathcal{O}_{\boldsymbol{\beta}}[\boldsymbol{\tau}]=\frac{C_{(-ib)}[-i(\boldsymbol{\tau}+\boldsymbol{\beta})]}{C_{(-ib)}[-i\boldsymbol{\beta}]}\,\mathrm{e}^{\odot N[\boldsymbol{\tau}]}.

We denote by 𝜷+,𝜷N\mathcal{F}^{N}_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}} the OPE family of N𝜷+,𝜷.N_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}.

Theorem 9.1.

Given two double background charges 𝛃±,𝛃0±\boldsymbol{\beta}^{\pm},\boldsymbol{\beta}_{0}^{\pm} with the neutrality conditions (NCb±),(\mathrm{NC}_{b^{\pm}}), 𝛃+,𝛃N\mathcal{F}^{N}_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}} is the image of 𝛃0+,𝛃0N\mathcal{F}^{N}_{\boldsymbol{\beta}_{0}^{+},\boldsymbol{\beta}_{0}^{-}} under the insertion of eN[𝛃+𝛃0+,𝛃𝛃0].\mathrm{e}^{\odot N[\boldsymbol{\beta}^{+}-\boldsymbol{\beta}_{0}^{+},\boldsymbol{\beta}^{-}-\boldsymbol{\beta}_{0}^{-}]}.

Proof.

By Wick’s calculus, we have

𝐄N𝜷0+,𝜷0(z,z)eN[𝜷+𝜷0+,𝜷𝜷0]=n𝜷0+,𝜷0(z,z)+𝐄N(z,z)N[𝜷+𝜷0+,𝜷𝜷0].\mathbf{E}\,N_{\boldsymbol{\beta}_{0}^{+},\boldsymbol{\beta}_{0}^{-}}(z,z^{\prime})\,\mathrm{e}^{\odot N[\boldsymbol{\beta}^{+}-\boldsymbol{\beta}_{0}^{+},\boldsymbol{\beta}^{-}-\boldsymbol{\beta}_{0}^{-}]}=n_{\boldsymbol{\beta}_{0}^{+},\boldsymbol{\beta}_{0}^{-}}(z,z^{\prime})+\mathbf{E}\,N(z,z^{\prime})N[\boldsymbol{\beta}^{+}-\boldsymbol{\beta}_{0}^{+},\boldsymbol{\beta}^{-}-\boldsymbol{\beta}_{0}^{-}].

In the \mathbb{H}-uniformization, it follows from (9.1) and (9.2) that

𝐄N(z,z)N[𝜷+𝜷0+,𝜷𝜷0]=n𝜷+,𝜷(z,z)n𝜷0+,𝜷0(z,z).\mathbf{E}\,N(z,z^{\prime})N[\boldsymbol{\beta}^{+}-\boldsymbol{\beta}_{0}^{+},\boldsymbol{\beta}^{-}-\boldsymbol{\beta}_{0}^{-}]=n_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}(z,z^{\prime})-n_{\boldsymbol{\beta}_{0}^{+},\boldsymbol{\beta}_{0}^{-}}(z,z^{\prime}).

Since both sides are scalars, the above identity holds for any chart. Thus we have

𝐄N𝜷0+,𝜷0(z,z)eN[𝜷+𝜷0+,𝜷𝜷0]=𝐄N𝜷+,𝜷(z,z).\mathbf{E}\,N_{\boldsymbol{\beta}_{0}^{+},\boldsymbol{\beta}_{0}^{-}}(z,z^{\prime})\,\mathrm{e}^{\odot N[\boldsymbol{\beta}^{+}-\boldsymbol{\beta}_{0}^{+},\boldsymbol{\beta}^{-}-\boldsymbol{\beta}_{0}^{-}]}=\mathbf{E}\,N_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}(z,z^{\prime}).

As in the Dirichlet case, we use Wick’s calculus again to derive

𝐄N𝜷0+,𝜷0(z1,z1)N𝜷0+,𝜷0(zn,zn)eN[𝜷+𝜷0+,𝜷𝜷0]=𝐄N𝜷+,𝜷(z1,z1)N𝜷+,𝜷(zn,zn).\mathbf{E}\,N_{\boldsymbol{\beta}_{0}^{+},\boldsymbol{\beta}_{0}^{-}}(z_{1},z_{1}^{\prime})\odot\cdots\odot N_{\boldsymbol{\beta}_{0}^{+},\boldsymbol{\beta}_{0}^{-}}(z_{n},z_{n}^{\prime})\,\mathrm{e}^{\odot N[\boldsymbol{\beta}^{+}-\boldsymbol{\beta}_{0}^{+},\boldsymbol{\beta}^{-}-\boldsymbol{\beta}_{0}^{-}]}=\mathbf{E}\,N_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}(z_{1},z_{1}^{\prime})\odot\cdots\odot N_{\boldsymbol{\beta}^{+},\boldsymbol{\beta}^{-}}(z_{n},z_{n}^{\prime}).

Differentiating with respect to zj,zjz_{j},z_{j}^{\prime}, we complete the proof. ∎

Example.

Given a marked boundary point qD,q\in\partial D, we consider a conformal map w:(D,q)(,).w:(D,q)\to(\mathbb{H},\infty). For 𝜷+=2bq\boldsymbol{\beta}^{+}=2b\cdot q and 𝜷=𝟎,\boldsymbol{\beta}^{-}=\boldsymbol{0},

N𝜷(z,z0)=N(z,z0)+2b(log|w(z)|log|w(z0)|).N_{\boldsymbol{\beta}}(z,z_{0})=N(z,z_{0})+2b(\log|w^{\prime}(z)|-\log|w^{\prime}(z_{0})|).

The function (z,z0)log|w(z)|log|w(z0)|(z,z_{0})\mapsto\log|w^{\prime}(z)|-\log|w^{\prime}(z_{0})| does not depend on the choice of w.w.

9.3. One-leg operators

We consider a simply-connected domain DD with two marked points pD,p\in\partial D, qDq\in\partial D in the chordal case, and qDq\in D in the radial case. For a symmetric background charge 𝜷\boldsymbol{\beta} on S=DdoubleS=D^{\mathrm{double}} with 𝜷(p)=a\boldsymbol{\beta}(p)=a and the neutrality condition (NCb),(\mathrm{NC}_{b}), the one-leg operator Λp\Lambda_{p} in the chordal case (the radial case, respectively) is defined by

ΛpΛ(p)=𝒪𝜷ˇ[𝝉],𝜷ˇ=𝜷𝝉\Lambda_{p}\equiv\Lambda(p)=\mathcal{O}_{\check{\boldsymbol{\beta}\,}}[\boldsymbol{\tau}],\qquad\check{\boldsymbol{\beta}\,}=\boldsymbol{\beta}-\boldsymbol{\tau}

where 𝝉=apaq\boldsymbol{\tau}=a\cdot p-a\cdot q in the chordal case (and 𝝉=ap12aq12aq\boldsymbol{\tau}=a\cdot p-\frac{1}{2}a\cdot q-\frac{1}{2}a\cdot q^{*} in the radial case, respectively). The insertion of Λp/𝐄Λp\Lambda_{p}/\mathbf{E}\,\Lambda_{p} is an operator

𝒳𝒳^\mathcal{X}\mapsto\widehat{\mathcal{X}}

on Fock space functionals/fields by the rules (5.20) and the formula in the chordal case

N^𝜷ˇ(z,z0)=N𝜷ˇ(z,z0)2a(log|w(z)|log|w(z0)|))\widehat{N}_{\check{\boldsymbol{\beta}\,}}(z,z_{0})=N_{\check{\boldsymbol{\beta}\,}}(z,z_{0})-2a\big{(}\log|w(z)|-\log|w(z_{0})|\big{)}\big{)}

where w:(D,p,q)(,0,)w:(D,p,q)\to(\mathbb{H},0,\infty) is a conformal map. In the radial case, the corresponding formula is given by

N^𝜷ˇ(z,z0)=N𝜷ˇ(z,z0)2a(log|1w(z)|log|1w(z0)|))+a(log|w(z)|log|w(z0)|))\widehat{N}_{\check{\boldsymbol{\beta}\,}}(z,z_{0})=N_{\check{\boldsymbol{\beta}\,}}(z,z_{0})-2a\big{(}\log|1-w(z)|-\log|1-w(z_{0})|\big{)}\big{)}+a\big{(}\log|w(z)|-\log|w(z_{0})|\big{)}\big{)}

where w:(D,p,q)(𝔻,1,0)w:(D,p,q)\to(\mathbb{D},1,0) is a conformal map.

We denote

𝐄^𝒳:=𝐄Λp𝒳𝐄Λp.\widehat{\mathbf{E}}\,\mathcal{X}:=\frac{\mathbf{E}\,\Lambda_{p}\mathcal{X}}{\mathbf{E}\,\Lambda_{p}}.

By Theorem 9.1, we have

𝐄^𝒳=𝐄𝒳^\widehat{\mathbf{E}}\,\mathcal{X}=\mathbf{E}\,\widehat{\mathcal{X}}

for any string 𝒳\mathcal{X} of fields in 𝜷ˇN.\mathcal{F}^{N}_{\check{\boldsymbol{\beta}\,}}.

We now show the level two degeneracy equations for Λ\Lambda if the parameters aa and bb satisfy 2a(a+b)=1.2a(a+b)=-1. We remark that aa and bb in (9.3) have such a relation.

Proposition 9.2.

Provided that 2a(a+b)=1,2a(a+b)=-1, we have

(9.4) T𝜷ˇΛ=12a22Λ.T_{\check{\boldsymbol{\beta}\,}}*\Lambda=-\frac{1}{2a^{2}}\partial^{2}\Lambda.
Proof.

Let a~=ia,b~=ib.\tilde{a}=-ia,\tilde{b}=-ib. In terms of the action of Virasoro generators and current generators, the one-leg operators are Virasoro primary holomorphic fields of conformal dimension λ=12a~2a~b~\lambda=\frac{1}{2}\tilde{a}^{2}-\tilde{a}\tilde{b} and current primary with charge q=a~.q=\tilde{a}. This implies the level two degeneracy equation (9.4) for the one-leg operator provided that 2a~(a~+b~)=1.2\tilde{a}(\tilde{a}+\tilde{b})=1.

We here present an alternate but direct proof using Wick’s calculus. For simplicity, we consider the standard chordal case 𝜷ˇ=2bq\check{\boldsymbol{\beta}\,}=2b\cdot q only. It is left to the reader as an exercise to prove (9.4) for general background charges. Since the difference is a differential, it suffices to show (9.4) in .\mathbb{H}. In the \mathbb{H}-uniformization we have

N𝜷ˇ=N,N𝜷ˇ=J𝜷ˇ=J,T𝜷ˇ=T+bJ,T=12JJ,Λ(z)=eaN+(z,q).N_{\check{\boldsymbol{\beta}\,}}=N,\quad\partial N_{\check{\boldsymbol{\beta}\,}}=J_{\check{\boldsymbol{\beta}\,}}=J,\quad T_{\check{\boldsymbol{\beta}\,}}=T+b\partial J,\quad T=-\frac{1}{2}J\odot J,\quad\Lambda(z)=\mathrm{e}^{\odot aN^{+}(z,q)}.

Let us first compute TΛ:T*\Lambda:

TΛ=TΛ+aJΛ.T*\Lambda=T\odot\Lambda+a\partial J\odot\Lambda.

Indeed, as ζz,\zeta\to z, we have

T(ζ)Λ(z)=12n=0ann!J2(ζ)(N+)n(z,q)=TΛ(z)+I+II+o(1),T(\zeta)\Lambda(z)=-\frac{1}{2}\sum_{n=0}^{\infty}\frac{a^{n}}{n!}J^{\odot 2}(\zeta)(N^{+})^{\odot n}(z,q)=T\odot\Lambda(z)+\mathrm{I}+\mathrm{II}+o(1),

where terms I\mathrm{I} and II\mathrm{II} come from 1 and 2 contractions, respectively. It follows from 𝐄J(ζ)N+(z,q)=1/(ζz)\mathbf{E}\,J(\zeta)N^{+}(z,q)=-1/(\zeta-z) that

I=aζzJ(ζ)N+(z,q).\mathrm{I}=\frac{a}{\zeta-z}J(\zeta)\odot N^{+}(z,q).

Thus its contribution to TΛT*\Lambda is aJΛ.a\partial J\odot\Lambda. On the other hand, we have

II=12a2(𝐄J(ζ)N+(z,q))2Λ(z)=12a2(ζz)2Λ(z).\mathrm{II}=-\frac{1}{2}a^{2}\big{(}\mathbf{E}\,J(\zeta)N^{+}(z,q)\big{)}^{2}\Lambda(z)=-\frac{1}{2}\frac{a^{2}}{(\zeta-z)^{2}}\Lambda(z).

It has no contribution to TΛ.T*\Lambda. To compute T𝜷ˇΛ,T_{\check{\boldsymbol{\beta}\,}}*\Lambda, we need to compute JΛ.\partial J*\Lambda. We have

J(ζ)Λ(z)=JΛ(z)+o(1).\partial J(\zeta)\Lambda(z)=\partial J\odot\Lambda(z)+o(1).

This implies that JΛ=JΛ\partial J*\Lambda=\partial J\odot\Lambda and

T𝜷ˇΛ=TΛ+(a+b)JΛ.T_{\check{\boldsymbol{\beta}\,}}*\Lambda=T\odot\Lambda+(a+b)\partial J\odot\Lambda.

Differentiating Λ(z)=eaN+(z,q),\Lambda(z)=\mathrm{e}^{\odot aN^{+}(z,q)}, we have

12a22Λ=12a2(aJΛ)=12a2(aJΛ+a2JJΛ)=12aJΛ+TΛ.-\frac{1}{2a^{2}}\partial^{2}\Lambda=-\frac{1}{2a^{2}}\partial(aJ\odot\Lambda)=-\frac{1}{2a^{2}}(a\partial J\odot\Lambda+a^{2}J\odot J\odot\Lambda)=-\frac{1}{2a}\partial J\odot\Lambda+T\odot\Lambda.

Now (9.4) follows provided that 2a(a+b)=1.2a(a+b)=-1.

9.4. Ward’s equations and BPZ-Cardy equations

We define the puncture operators by 𝒫𝜷:=C(ib)[i𝜷].\mathcal{P}_{\boldsymbol{\beta}}:=C_{(-ib)}[-i\boldsymbol{\beta}]. As in the Dirichlet case, Ward’s equations hold for the extended OPE family 𝜷N\mathcal{F}^{N}_{\boldsymbol{\beta}} of N𝜷.N_{\boldsymbol{\beta}}. Since its proof is similar to the Dirichlet case, we leave it to the reader as an exercise.

Theorem 9.3 (Ward’s equations).

Let Y,X1,,Xn𝛃NY,X_{1},\cdots,X_{n}\in\mathcal{F}^{N}_{\boldsymbol{\beta}} and let XX be the tensor product of XjX_{j}’s. Then

𝐄T𝜷(z)X=𝐄𝒫𝜷1kz+𝒫𝜷X+𝐄𝒫𝜷1kz¯𝒫𝜷X,\mathbf{E}\,T_{\boldsymbol{\beta}}(z)\,X=\mathbf{E}\,\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{+}_{k_{z}}\mathcal{P}_{\boldsymbol{\beta}}X+\mathbf{E}\,\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{-}_{k_{\bar{z}}}\mathcal{P}_{\boldsymbol{\beta}}X,

where all fields are evaluated in the identity chart of \mathbb{H} and

2z2𝐄T𝜷(z)X=𝐄𝒫𝜷1vz+𝒫𝜷X+𝐄𝒫𝜷1vz𝒫𝜷X,2z^{2}\mathbf{E}\,T_{\boldsymbol{\beta}}(z)\,X=\mathbf{E}\,\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{+}_{v_{z}}\mathcal{P}_{\boldsymbol{\beta}}X+\mathbf{E}\,\mathcal{P}_{\boldsymbol{\beta}}^{-1}\mathcal{L}^{-}_{v_{z^{*}}}\mathcal{P}_{\boldsymbol{\beta}}X,

where all fields are evaluated in the identity chart of 𝔻.\mathbb{D}.

For a symmetric background charge 𝜷\boldsymbol{\beta} with the neutrality condition (NCb)(\mathrm{NC}_{b}) and a specific charge aa at a marked boundary point pD,p\in\partial D, we define the backward SLE partition function Z𝜷Z_{\boldsymbol{\beta}} by

Z𝜷:=|𝒫𝜷|=|C(ib)[i𝜷]|Z_{\boldsymbol{\beta}}:=|\mathcal{P}_{\boldsymbol{\beta}}|=|C_{(-ib)}[-i\boldsymbol{\beta}]|

and the effective one-leg operators by Λpeff:=𝒫𝜷Λp.\Lambda_{p}^{\mathrm{eff}}:=\mathcal{P}_{\boldsymbol{\beta}}\,\Lambda_{p}. Denote

(9.5) Rξ𝐄^ξ𝒳:=𝐄Λξ𝒳𝐄Λξ=𝐄Λξeff𝒳𝐄Λξeff.R_{\xi}\equiv\widehat{\mathbf{E}}_{\xi}\,\mathcal{X}:=\frac{\mathbf{E}\,\Lambda_{\xi}\mathcal{X}}{\mathbf{E}\,\Lambda_{\xi}}=\frac{\mathbf{E}\,\Lambda_{\xi}^{\mathrm{eff}}\mathcal{X}}{\mathbf{E}\,\Lambda_{\xi}^{\mathrm{eff}}}.

Let ξD\xi\in\partial D and 𝜷ξ=𝜷ap+aξ.\boldsymbol{\beta}_{\xi}=\boldsymbol{\beta}-a\cdot p+a\cdot\xi. Let XX be the tensor product X=X1(z1)Xn(zn)X=X_{1}(z_{1})\cdots X_{n}(z_{n}) of fields XjX_{j} in 𝜷ˇ\mathcal{F}_{\check{\boldsymbol{\beta}\,}} (zjD).(z_{j}\in D). Combining Proposition 9.2 with Theorem 9.3, we obtain the BPZ equations:

(9.6) 12a2ξ2𝐄ΛξeffX=𝐄ˇkξΛξeffX,kξ(z):=1ξz-\frac{1}{2a^{2}}\partial_{\xi}^{2}\mathbf{E}\Lambda_{\xi}^{\mathrm{eff}}X=\mathbf{E}\,\check{\mathcal{L}}_{k_{\xi}}\Lambda_{\xi}^{\mathrm{eff}}X,\qquad k_{\xi}(z):=\frac{1}{\xi-z}

in the \mathbb{H}-uniformization. In particular, the partition function ZξZ𝜷ξ:=|𝐄Λξeff|=|C(ib)[i𝜷ξ]|Z_{\xi}\equiv Z_{\boldsymbol{\beta}_{\xi}}:=\big{|}\mathbf{E}\,\Lambda^{\mathrm{eff}}_{\xi}\big{|}=\big{|}C_{(-ib)}[-i\boldsymbol{\beta}_{\xi}]\big{|} satisfies the null vector equation

(9.7) 12a2ξ2Zξ=ˇkξZξ-\frac{1}{2a^{2}}\partial_{\xi}^{2}Z_{\xi}=\check{\mathcal{L}}_{k_{\xi}}Z_{\xi}

in the \mathbb{H}-uniformization. Here ξ=+¯\partial_{\xi}=\partial+\bar{\partial} is the operator of differentiation with respect to the real variable ξ\xi and and ˇkξ\check{\mathcal{L}}_{k_{\xi}} is taken over the finite notes of XX and supp𝜷ξ{ξ}.\mathrm{supp}\,\boldsymbol{\beta}_{\xi}\setminus\{\xi\}. We obtain the following form of BPZ-Cardy equations in the chordal case. The radial case can be derived in a similar way.

Theorem 9.4 (BPZ-Cardy equations).

Suppose that the parameters aa and bb satisfy

2a(a+b)=1.2a(a+b)=-1.

If X=X1(z1)Xn(zn)X=X_{1}(z_{1})\cdots X_{n}(z_{n}) with Xj𝛃ˇN,X_{j}\in\mathcal{F}^{N}_{\check{\boldsymbol{\beta}\,}}, then we have

(9.8) 12a2(ξ2𝐄^ξX+2(ξlogZξ)ξ𝐄^ξX)=ˇkξ𝐄^ξX-\frac{1}{2a^{2}}\big{(}\partial_{\xi}^{2}\widehat{\mathbf{E}}_{\xi}\,X+2(\partial_{\xi}\log Z_{\xi})\partial_{\xi}\widehat{\mathbf{E}}_{\xi}\,X\big{)}=\check{\mathcal{L}}_{k_{\xi}}\widehat{\mathbf{E}}_{\xi}\,X

in the identity chart of .\mathbb{H}. In the (𝔻,0)(\mathbb{D},0)-uniformization we have

(9.9) κ(12θ2+(θlogZζ)θ)𝐄^ζX=ˇvζ𝐄^ζX,ζ=eiθ.\kappa\Big{(}\frac{1}{2}\partial_{\theta}^{2}+(\partial_{\theta}\log Z_{\zeta}\big{)}\partial_{\theta}\Big{)}\widehat{\mathbf{E}}_{\zeta}X=\check{\mathcal{L}}_{v_{\zeta}}\widehat{\mathbf{E}}_{\zeta}X,\qquad\zeta=\mathrm{e}^{i\theta}.

9.5. Connection to backward SLE theory

We now prove Theorem 1.6 and present Sheffield’s observables. By definition, the backward radial (see (1.11) for the chordal case) SLE[𝜷]\mathrm{SLE}[\boldsymbol{\beta}] map ftf_{t} from 𝔻\mathbb{D} satisfies the equation

(9.10) tft(z)=ft(z)ζt+ft(z)ζtft(z),ζt=eiθt\partial_{t}f_{t}(z)=-f_{t}(z)\frac{\zeta_{t}+f_{t}(z)}{\zeta_{t}-f_{t}(z)},\qquad\zeta_{t}=\mathrm{e}^{i\theta_{t}}

driven by the real process θt:\theta_{t}:

dθt=κdBt+λ(t)dt,λ(t)=(λft1),λ=κθlogZ𝜷ζ,\mathrm{d}\theta_{t}=\sqrt{\kappa}\,\mathrm{d}B_{t}+\lambda(t)\,\mathrm{d}t,\quad\lambda(t)=(\lambda\,\|\,f_{t}^{-1}),\quad\lambda=\kappa\,\partial_{\theta}\log Z_{\boldsymbol{\beta}_{\zeta}},

where the partition function Z𝜷ζZ_{\boldsymbol{\beta}_{\zeta}} is given by Z𝜷ζ=|C(ib)[i𝜷ζ]|,Z_{\boldsymbol{\beta}_{\zeta}}=\big{|}C_{(-ib)}[-i\boldsymbol{\beta}_{\zeta}]\big{|}, 𝜷ζ=𝜷+aζap.\boldsymbol{\beta}_{\zeta}=\boldsymbol{\beta}+a\cdot\zeta-a\cdot p.

Proof of Theorem 1.6.

We first consider the chordal case. For 𝒳=X1(z1)Xn(zn),Xj𝜷ˇ(D),\mathcal{X}=X_{1}(z_{1})\cdots X_{n}(z_{n}),X_{j}\in\mathcal{F}_{\check{\boldsymbol{\beta}\,}}(D), denote

Rξ(z1,,zn)𝐄^ξ[X1(z1)Xn(zn)].R_{\xi}(z_{1},\cdots,z_{n})\equiv\widehat{\mathbf{E}}_{\xi}[X_{1}(z_{1})\cdots X_{n}(z_{n})].

We want to show that

Mt=m(ξt,t),m(ξ,t)=(Rξft1)M_{t}=m(\xi_{t},t),\quad m(\xi,t)=(R_{\xi}\,\|\,f_{t}^{-1})

is a local martingale on backward chordal SLE probability space. By Itô’s formula, we compute dMt\mathrm{d}M_{t} as follows:

dMt\displaystyle\mathrm{d}M_{t} =κξ|ξ=ξtm(ξ,t)dBt+κξ|ξ=ξtm(ξ,t)ξ|ξ=ξt(logZξft1)\displaystyle=\sqrt{\kappa}\,\partial_{\xi}\big{|}_{\xi=\xi_{t}}m(\xi,t)\,\mathrm{d}B_{t}+\kappa\,\partial_{\xi}\big{|}_{\xi=\xi_{t}}m(\xi,t)\,\partial_{\xi}\big{|}_{\xi=\xi_{t}}(\log Z_{\xi}\,\|\,f_{t}^{-1})
+κ2ξ2|ξ=ξtm(ξ,t)dt+dds|s=0(Rξtft+s1)dt.\displaystyle+\frac{\kappa}{2}\,\partial_{\xi}^{2}\big{|}_{\xi=\xi_{t}}m(\xi,t)\,\mathrm{d}t+\frac{\mathrm{d}}{\mathrm{d}s}\Big{|}_{s=0}\big{(}R_{\xi_{t}}\,\|\,f_{t+s}^{-1}\big{)}\,\mathrm{d}t.

Using the same argument in the proof of Theorem 1.5, the last term can be rewritten as

dds|s=0(Rξtft+s1)dt=2(ˇkξtRξtft1).\frac{\mathrm{d}}{\mathrm{d}s}\Big{|}_{s=0}\big{(}R_{\xi_{t}}\,\|\,f_{t+s}^{-1}\big{)}\,\mathrm{d}t=2\big{(}\check{\mathcal{L}}_{k_{\xi_{t}}}R_{\xi_{t}}\,\|\,f_{t}^{-1}\big{)}.

Thus we find the drift term of dMt\mathrm{d}M_{t} as

(κ2ξ2|ξ=ξtm(ξ,t)+κξ|ξ=ξtm(ξ,t)ξ|ξ=ξt(Zξft1)(Zξft1)+2(ˇkξtRξtft1))dt=0\Big{(}\frac{\kappa}{2}\,\partial_{\xi}^{2}\big{|}_{\xi=\xi_{t}}m(\xi,t)+\kappa\,\partial_{\xi}\big{|}_{\xi=\xi_{t}}m(\xi,t)\frac{\partial_{\xi}|_{\xi=\xi_{t}}(Z_{\xi}\,\|\,f_{t}^{-1})}{(Z_{\xi}\,\|\,f_{t}^{-1})}+2\big{(}\check{\mathcal{L}}_{k_{\xi_{t}}}R_{\xi_{t}}\,\|\,f_{t}^{-1}\big{)}\Big{)}\,\mathrm{d}t=0

employing the BPZ-Cardy equations (9.8).

Next, we consider the radial case. For 𝒳=X1(z1)Xn(zn),Xj𝜷ˇ(D),\mathcal{X}=X_{1}(z_{1})\cdots X_{n}(z_{n}),X_{j}\in\mathcal{F}_{\check{\boldsymbol{\beta}\,}}(D), denote

Rζ(z1,,zn)𝐄^ζ[X1(z1)Xn(zn)].R_{\zeta}(z_{1},\cdots,z_{n})\equiv\widehat{\mathbf{E}}_{\zeta}[X_{1}(z_{1})\cdots X_{n}(z_{n})].

As in the chordal case, the process

Mt=m(ζt,t),m(ζ,t)=(Rζft1)M_{t}=m(\zeta_{t},t),\qquad m(\zeta,t)=\big{(}R_{\zeta}\,\|\,f_{t}^{-1}\big{)}

is a local martingale by Itô’s formula and the BPZ-Cardy equations (9.9). Indeed, we find the drift term of dMt\mathrm{d}M_{t} as

(κ2θ2+κθ(Zζft1)(Zζft1)θ)|ζ=ζtm(ζ,t)dt(ˇvζtRζtft1)dt=0.\Big{(}\frac{\kappa}{2}\partial_{\theta}^{2}+\kappa\frac{\partial_{\theta}(Z_{\zeta}\,\|\,f_{t}^{-1})}{(Z_{\zeta}\,\|\,f_{t}^{-1})}\partial_{\theta}\Big{)}\Big{|}_{\zeta=\zeta_{t}}~{}m(\zeta,t)\,\mathrm{d}t\\ -\ \big{(}\check{\mathcal{L}}_{v_{\zeta_{t}}}R_{\zeta_{t}}\,\|\,f_{t}^{-1}\big{)}\mathrm{d}t=0.

Example (Sheffield’s observables).

In the standard chordal case with 𝜷=ap+(2ba)q,\boldsymbol{\beta}=a\cdot p+(2b-a)\cdot q, the 1-point functions of the bosonic fields

nt(z)\displaystyle n_{t}(z) =𝐄[N𝜷(z)ft1]=2alog|ft(z)ξt|+2blog|ft(z)|\displaystyle=\mathbf{E}[N_{\boldsymbol{\beta}}(z)\,\|\,f_{t}^{-1}]=-2a\log|f_{t}(z)-\xi_{t}|+2b\log|f_{t}^{\prime}(z)|
=2alog|z|+220tRe1fs(z)ξsdBs\displaystyle=-2a\log|z|+2\sqrt{2}\int_{0}^{t}\mathrm{Re}\,\frac{1}{f_{s}(z)-\xi_{s}}\,\mathrm{d}B_{s}

were introduced as backward SLE martingale-observables in [34]. Due to the following special case of Hadamard’s variation formula

(9.11) dGN(ft(z1),ft(z2))=4Re1ft(z1)ξtRe1ft(z2)ξtdt=12dn𝜷(z1),n𝜷(z2)t,\mathrm{d}G^{N}(f_{t}(z_{1}),f_{t}(z_{2}))=-4\,\mathrm{Re}\,\frac{1}{f_{t}(z_{1})-\xi_{t}}\,\mathrm{Re}\,\frac{1}{f_{t}(z_{2})-\xi_{t}}\,\mathrm{d}t=-\frac{1}{2}\,\mathrm{d}\langle n_{\boldsymbol{\beta}}(z_{1}),n_{\boldsymbol{\beta}}(z_{2})\rangle_{t},

the formal 22-point functions

𝐄[N𝜷(z1)N𝜷(z2)]=2GN(z1,z2)+n𝜷(z1)n𝜷(z2)\mathbf{E}[N_{\boldsymbol{\beta}}(z_{1})N_{\boldsymbol{\beta}}(z_{2})]=2G^{N}(z_{1},z_{2})+n_{\boldsymbol{\beta}}(z_{1})n_{\boldsymbol{\beta}}(z_{2})

are martingale-observables for backward SLE. Sheffield used (9.11) to construct a coupling of backward SLE and the Gaussian free field with Neumann boundary condition, see [34].

Acknowledgements

We would like to thank Tom Alberts, and Sung-Soo Byun for careful reading and much-appreciated help improving this manuscript.

References

  • [1] Tom Alberts, Nam-Gyu Kang, and Nikolai G. Makarov. Conformal field theory for multiple SLEs. in preparation.
  • [2] Tom Alberts, Nam-Gyu Kang, and Nikolai G. Makarov. Pole dynamics and an integral of motion for multiple SLE(0). 2020. arXiv:2011.05714.
  • [3] Tom Alberts, Michael J. Kozdron, and Gregory F. Lawler. The Green function for the radial Schramm-Loewner evolution. J. Phys. A, 45(49):494015, 17, 2012.
  • [4] Michel Bauer and Denis Bernard. Conformal field theories of stochastic Loewner evolutions. Comm. Math. Phys., 239(3):493–521, 2003. arXiv:hep-th/0210015.
  • [5] Michel Bauer and Denis Bernard. CFTs of SLEs: the radial case. Phys. Lett. B, 583(3-4):324–330, 2004.
  • [6] Vincent Beffara. The dimension of the SLE curves. Ann. Probab., 36(4):1421–1452, 2008.
  • [7] A. A. Belavin, A. M. Polyakov, and A. B. Zamolodchikov. Infinite conformal symmetry in two-dimensional quantum field theory. Nuclear Phys. B, 241(2):333–380, 1984.
  • [8] Sung-Soo Byun, Nam-Gyu Kang, and Hee-Joon Tak. Conformal field theory for annulus SLE: partition functions and martingale-observables. 2018. arXiv:1806.03638v2.
  • [9] John Cardy. Calogero-Sutherland model and bulk-boundary correlations in conformal field theory. Phys. Lett. B, 582(1-2):121–126, 2004.
  • [10] John Cardy. SLE(kappa,rho) and Conformal Field Theory. 2004. arXiv:math-ph/0412033.
  • [11] Philippe Di Francesco, Pierre Mathieu, and David Sénéchal. Conformal field theory. Graduate Texts in Contemporary Physics. Springer-Verlag, New York, 1997.
  • [12] B. Doyon and J. Cardy. Calogero-Sutherland eigenfunctions with mixed boundary conditions and conformal field theory correlators. J. Phys. A, 40(10):2509–2540, 2007.
  • [13] Julien Dubédat. SLE(κ,ρ){\rm SLE}(\kappa,\rho) martingales and duality. Ann. Probab., 33(1):223–243, 2005.
  • [14] Julien Dubédat. SLE and the free field: partition functions and couplings. J. Amer. Math. Soc., 22(4):995–1054, 2009.
  • [15] Roland Friedrich and Wendelin Werner. Conformal restriction, highest-weight representations and SLE. Comm. Math. Phys., 243(1):105–122, 2003.
  • [16] V. G. Kac and A. K. Raina. Bombay lectures on highest weight representations of infinite-dimensional Lie algebras, volume 2 of Advanced Series in Mathematical Physics. World Scientific Publishing Co. Inc., Teaneck, NJ, 1987.
  • [17] Nam-Gyu Kang and Nikolai G. Makarov. Calculus of conformal fields on a compact Riemann surface. arXiv:1708.07361.
  • [18] Nam-Gyu Kang and Nikolai G. Makarov. Gaussian free field and conformal field theory. Astérisque, (353):viii+136, 2013.
  • [19] Nam-Gyu Kang and Hee-Joon Tak. Conformal field theory of dipolar SLE with the Dirichlet boundary condition. Anal. Math. Phys., 3(4):333–373, 2013.
  • [20] Kalle Kytölä. On conformal field theory of SLE(κ,ρ){\rm SLE}(\kappa,\rho). J. Stat. Phys., 123(6):1169–1181, 2006.
  • [21] Gregory F. Lawler. Conformally invariant processes in the plane, volume 114 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2005.
  • [22] Gregory F. Lawler. Partition functions, loop measure, and versions of SLE. J. Stat. Phys., 134(5-6):813–837, 2009.
  • [23] Gregory F. Lawler, Oded Schramm, and Wendelin Werner. The dimension of the planar Brownian frontier is 4/34/3. Math. Res. Lett., 8(4):401–411, 2001.
  • [24] Gregory F. Lawler, Oded Schramm, and Wendelin Werner. Values of Brownian intersection exponents. II. Plane exponents. Acta Math., 187(2):275–308, 2001.
  • [25] Gregory F. Lawler, Oded Schramm, and Wendelin Werner. Conformal invariance of planar loop-erased random walks and uniform spanning trees. Ann. Probab., 32(1B):939–995, 2004.
  • [26] Jason Miller and Scott Sheffield. Imaginary geometry I: interacting SLEs. Probab. Theory Related Fields, 164(3-4):553–705, 2016.
  • [27] Jason Miller and Scott Sheffield. Imaginary geometry IV: interior rays, whole-plane reversibility, and space-filling trees. Probab. Theory Related Fields, 169(3-4):729–869, 2017.
  • [28] Eveliina Peltola and Yilin Wang. Large deviations of multichordal SLE0+, real rational functions, and zeta-regularized determinants of Laplacians. arXiv:2006.08574, to appear in J. Eur. Math. Soc.
  • [29] I. Rushkin, E. Bettelheim, I. A. Gruzberg, and P. Wiegmann. Critical curves in conformally invariant statistical systems. J. Phys. A, 40(9):2165–2195, 2007.
  • [30] Oded Schramm. Scaling limits of loop-erased random walks and uniform spanning trees. Israel J. Math., 118:221–288, 2000.
  • [31] Oded Schramm and Scott Sheffield. Contour lines of the two-dimensional discrete Gaussian free field. Acta Math., 202(1):21–137, 2009.
  • [32] Oded Schramm and Scott Sheffield. A contour line of the continuum Gaussian free field. Probab. Theory Related Fields, 157(1-2):47–80, 2013.
  • [33] Oded Schramm and David B. Wilson. SLE coordinate changes. New York J. Math., 11:659–669 (electronic), 2005.
  • [34] Scott Sheffield. Conformal weldings of random surfaces: SLE and the quantum gravity zipper. Ann. Probab., 44(5):3474–3545, 2016.
  • [35] Stanislav Smirnov. Critical percolation in the plane: conformal invariance, Cardy’s formula, scaling limits. C. R. Acad. Sci. Paris Sér. I Math., 333(3):239–244, 2001.
  • [36] Stanislav Smirnov. Conformal invariance in random cluster models. I. Holomorphic fermions in the Ising model. Ann. of Math. (2), 172(2):1435–1467, 2010.
  • [37] Dapeng Zhan. Duality of chordal SLE. Invent. Math., 174(2):309–353, 2008.