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

institutetext: Kavli Institute for Theoretical Sciences, University of Chinese Academy of Sciences,
Beijing 100190, China

Shape dependence of mutual information in the OPE limit: linear responses

Liangyu Chen    and Huajia Wang [email protected] [email protected]
Abstract

Mutual information serves as an important measure of correlation between subsystem components. In the framework of quantum field theories (QFTs) they have better regulated UV behavior than entanglement entropy, and thus provide more direct access to universal aspects of entanglement structures. In this paper, we study the linear responses under shape deformation of the mutual information in the conformal field theory (CFT) vacuum between two spheres of radius RR separated by large distance LRL\gg R or conformally equivalent configurations. Our calculations make use of the previous OPE results for mutual information Faulkner2016Aug and the associated modular Hamiltonian Faulkner2021Aug . In particular, we apply the entanglement first law to compute the linear responses of mutual information under shape deformation on one of the spheres. We find that the linear responses exhibit a high degree of universality for a selected class of OPE contributions. We demonstrate that there is a “little group” of symmetries associated with the set-up. Our result implies that the spherical mutual information is extremal over shape deformations of non-zero modes under the symmetry group.

1 Introduction

Entanglement structure has played a key role in probing deep aspects of quantum field theories (QFTs) not accessible using conventional tools and techniques. For example, it can provide the order parameter for topological phase Kitaev2005Mar ; Xiao-GangMar ; through the famous Ryu-Takayanagi formula Ryu2006 and its subsequent generalizations Hubeny2007 , it encodes information regarding bulk space-time geometry in holography Maldacena1999 ; Witten1998AntideSS ; GUBSER1998105 ; it can be used to characterize universal monotonous properties of renormalization group flow Myers2011Jan ; Casini2011May ; Casini:2015woa ; and it is also crucial for revealing connection between information theory and energy conditions HWang2016Sep ; HWang2019Sep ; Faulkner1804 , etc.

In principle, one can define various measures of the underlying entanglement structure. An important one is the entanglement entropy between subsystem AA and its complement A¯\bar{A}. For a global pure state |ψ|\psi\rangle, the entanglement entropy is defined by:

SAψtrAρAψlnρAψ,ρAψ=trA¯|ψψ|S^{\psi}_{A}\equiv-\text{tr}_{A}\rho^{\psi}_{A}\ln{\rho^{\psi}_{A}},\;\;\rho^{\psi}_{A}=\text{tr}_{\bar{A}}|\psi\rangle\langle\psi| (1)

This definition relies on the notion of reduced density matrix ρAψ\rho^{\psi}_{A}, which only makes clear sense if the global Hilbert space factorizes into tensor product =AA¯\mathcal{H}=\mathcal{H}_{A}\otimes\mathcal{H}_{\bar{A}}. This is a convenient but usually problematic assumption in quantum field theories, for example due to gauge constraints imposed at the entangling boundary A\partial A William2012 ; Casini2014Apr ; Lin:2018 . Another aspect associated with A\partial A in quantum field theories is the ubiquitous ultra-violet (UV) divergences that arise when computing entanglement entropies. Roughly speaking, they come from short-range entanglements across A\partial A that are only regularized by the UV cut-off of the theory. For example, in CFT the divergences take the form:

S=cd2Rd2δd2+cd4Rd4δd4+{+c1,d= odd+c0log(Rδ),d= even}S=c_{d-2}\frac{R^{d-2}}{\delta^{d-2}}+c_{d-4}\frac{R^{d-4}}{\delta^{d-4}}+...\begin{Bmatrix}&+&c_{-1},\;\;\;d=\text{ odd}\\ &+&c_{0}\log{\left(\frac{R}{\delta}\right)},\;d=\text{ even}\end{Bmatrix} (2)

where R,δR,\delta is the scale of entangling region and UV cutoff respectively, cd2,cd4,c_{d-2},c_{d-4},... are coefficients that depend on the geometry of entangling regions, and c1c_{-1} and c0c_{0} are universal. As a result, additional efforts are required in order to extract universal aspects of entanglement structures from the result of entanglement entropy. In order to do this, one usually construct combinations of entanglement quantities whose UV dependences cancel out. In practice, such cancellations can be arranged in a number of ways. For example, when extracting the order parameters for topological phases in 2+1 dimensional gapped systems, one computes the so-called topological entanglement entropy Kitaev2005Mar ; Xiao-GangMar by adding and subtracting vacuum entanglement entropies associated with different regions:

StopoSA+SB+SCSABSBCSAC+SABCS_{\text{topo}}\equiv S_{A}+S_{B}+S_{C}-S_{AB}-S_{BC}-S_{AC}+S_{ABC} (3)

out of which the shape dependence of these regions cancel out completely and leaving only topological contributions. Fixing a particular entangling region AA, one can also arrange the cancellation between states, for example when computing the so-called relative entropy between two global states ψ\psi and σ\sigma:

SA(ψ|σ)trρAψlnρAψtrρAψlnρAσS_{A}\left(\psi|\sigma\right)\equiv\text{tr}\rho^{\psi}_{A}\ln{\rho^{\psi}_{A}}-\text{tr}\rho^{\psi}_{A}\ln{\rho^{\sigma}_{A}} (4)

The relative entropy SA(ψ|σ)S_{A}\left(\psi|\sigma\right) provides a measure of distinguishability between the global states ψ\psi and σ\sigma based on what one can access in the subsystem AA, i.e. between ρAψ\rho^{\psi}_{A} and ρAσ\rho^{\sigma}_{A}. It satisfies several constraints such as positivity and monotonicity, and when applied to specific context they can sometimes reveal important physics such as the emergence of Einstein’s equation in AdS/CFT Faulkner2014Mar ; Faulkner2015May ; Faulkner2017Aug ; Lewkowycz2018 or the validity of the averaged null energy condition (ANEC) in QFTs HWang2016Sep , etc.

Mutual information is one of the simplest combinations among these entanglement quantities. For disjoint entangling regions AA and BB, one can define the mutual information IA,BI_{A,B} between them as:

IA,BSABSASBI_{A,B}\equiv S_{A\cup B}-S_{A}-S_{B} (5)

where we have omitted the dependence on the state ψ\psi in these quantities, since in this paper shall focus on the vacuum state |ψ=|Ω|\psi\rangle=|\Omega\rangle from now on. The UV divergences associated with short-distance entanglements near (AB)\partial\left(A\cup B\right) are cancelled in IA,BI_{A,B}, thus making it a universal quantity. Roughly speaking IABI_{AB} measures the correlation between the subsystems AA and BB. When we view ABA\cup B together as a system in the mixed state ρAB\rho_{AB}, IA,BI_{A,B} receives contributions from both classical and quantum correlations. In attempts to refine the distinction between classical and quantum correlations in mixed states, other measures have been proposed which include entanglement negativity Eisert1999 ; Vidal2002 ; Plenio2005 . Being a universal quantity, the mutual information behaves in ways that align more with intuitions. For example due to strong subadditivity (SSA) for entanglement entropies, the mutual information like relative entropies also satisfy monotonicity, i.e.:

IA,BIA~,BI_{A,B}\leq I_{\tilde{A},B} (6)

for AA~A\subseteq\tilde{A}, while the entanglement entropies in general do not satisfy such constraints by themselves. On the other hand, the mutual information is less constrained by symmetries and thus encode more detailed information about the underlying theories. For example, while the single-interval vacuum entanglement entropies in 2d CFT ares constrained by symmetries to only depend on the central charges cc, the mutual information between single-intervals depend on the full operator-spectrum of the CFTs.

Apart from the choice of entanglement measures, shape dependence on the entangling region can also reveal important information regarding entanglement structure. For example, in the presence of corner or cusp the entanglement entropy receives log divergent contributions whose coefficients on the other hand contain universal information Fradkin2006 ; Casini2007Dec ; Hirata2007 . For regions with smooth boundaries, more generally one can study shape dependence by computing perturbation theory under shape deformation Rosenhaus2014Dec ; Rosenhaus2015Feb , usually about symmetric shapes i.e. spheres, half-planes etc. For spheres in CFTs, computing the first-order response indicates that its entanglement entropy is extremal under symmetry-breaking deformations Mezei2015Feb ; Mezei2015Feb02 ; while it being perturbatively minimal was verified by the second-order response which also revealed universal non-local contributions known as the entanglement density Faulkner2016Apr ; the minimization was later proven beyond perturbation theory Bueno2021 . Sometimes by probing shape dependence of carefully chosen entanglement quantities, one discover surprising connections between entanglement behaviors and other aspects of QFTs, such as the irreversibility of RG flows and the validity of energy conditions, manifested through analyzing shape dependence of entanglement entropies (appropriately regularized) and relative entropies.

It is therefore natural to expect that the shape dependence of mutual information can also reflect important physics. For example, it has been shown that the shape dependence of IA,BI_{A,B} satisfies constraints on the null cone of AA which in the large separation limit (see 8 below) between AA and BB can give rise to the unitarity bound Casini2021Sep . However, explicit results for shape dependence of mutual information using the general perturbative approach have been missing so far. One reason for this is that we do not have configurations with non-trivial mutual information that are sufficiently symmetric. In the case of a single sphere AA the vacuum entanglement structure in CFT is fixed by symmetry. In particular the modular Hamiltonian which encodes the complete entanglement data is known explicitly Casini2011May :

H^A=2πA𝑑x(R2|x2|2R)T^tt(x),ρA=eH^A\hat{H}_{A}=2\pi\int_{A}dx\left(\frac{R^{2}-|x^{2}|}{2R}\right)\hat{T}_{tt}(x),\;\;\rho_{A}=e^{-\hat{H}_{A}} (7)

Therefore it makes sense to do perturbation theory about it as in Mezei2015Feb ; Mezei2015Feb02 ; Faulkner2016Apr . On the other hand, the mutual information IA,BI_{A,B} between two spheres AA and BB in CFT vacuum is not known in general form, despite being the most symmetric configuration with non-trivial mutual information. As mentioned before it is sensitive to the details (e.g. operator spectrum) of the theory. Explicit results can be obtained by restricting to special theories, e.g. Dirac fermions in two-dimensions Casini2005Jul ; Casini2009Mar ; Casini2009Sep , in which the shape dependence on entangling region is degenerate; or by restricting to special limits in general CFTs, e.g. of large separation LL between AA and BB (see Fig 1), where IABI_{AB} admits an OPE-like expansion Cardy2013 ; Faulkner2016Aug ; ChenBin2017 ; Long2016 :

IA,B=𝒩ΔπΓ(2Δ+1)4Γ(2Δ+32)(RARBL2)2Δ+I_{A,B}=\mathcal{N}_{\Delta}\frac{\sqrt{\pi}\Gamma(2\Delta+1)}{4\Gamma\left(2\Delta+\frac{3}{2}\right)}\left(\frac{R_{A}R_{B}}{L^{2}}\right)^{2\Delta}+\ldots (8)

where Δ\Delta denotes the conformal dimension of an internal scalar primary operator that “carries” the correlation between AA and BB in some loose sense. The OPE result (8) is valid in any dimensional CFTs, and therefore provides a good starting point to develop perturbative expansion in shape deformations.

Refer to caption
Figure 1: Two conformally equivalent configurations for the mutual information. The left: two spherical subregions of radius RR separated by a distance L0RL_{0}\gg R. The right: two spherical subregions of radius RAR_{A} and RBR_{B} separated with distance LRARBL\sim R_{A}\gg R_{B}.

In this paper, we study the shape dependence of mutual information perturbatively about spheres. We will focus on the leading order contribution in the OPE (8). More specifically, we will compute the first-order linear response under shape deformation about one of the spheres AA. The OPE expansion (8) is conformally invariant, i.e. depending only on the conformal-ratio:

ρ=2RA2RBL2RA2RB21\rho=\frac{2R_{A}2R_{B}}{L^{2}-R_{A}^{2}-R_{B}^{2}}\ll 1 (9)

where RA,BR_{A,B} are the radius of the spheres A,BA,B and LL is the distance between them. Therefore, we have the freedom to work in a conformal frame where RALRBR_{A}\sim L\gg R_{B}, see Fig. 1, so that the fine-details in the shape-response of IA,BI_{A,B} on AA are optimally pronounced. A key input into our calculation is the modular Hamiltonian version of the OPE expansion (8) derived in Faulkner2021Aug . This is because at the first order, the shape perturbation theory can be implemented in the context of the entanglement first law, which requires knowledge about the unperturbed modular Hamiltonian.

The paper is organized as follows. In section 2, we recall some basic ingredients regarding shape perturbation theory and its application through the entanglement first law for computing linear responses. In section 3, we summarize previous results for mutual information and the corresponding modular Hamiltonians in the OPE limit, which our calculations will be based upon. In section 4, we carry out the main computation and obtain results for linear response of mutual information under shape deformation on one of the spheres. During this we identify some subtleties associated with the integral expression of modular Hamiltonian in terms of local operators. In section 5, we explore the implications of our results from a symmetry point of view, and discuss the extremality property under shape deformations for mutual information on spheres. In section 6, we conclude the paper with some discussions and suggestions for future directions.

2 Shape perturbation theory and the entanglement first law

In this section, we quickly recall the necessary ingredients needed for our computations. More details can be found in Rosenhaus2014Dec ; Rosenhaus2015Feb ; Mezei2015Feb ; Mezei2015Feb02 ; Faulkner2016Apr .

2.1 Shape perturbation theory

We are only interested in the vacuum states, whose wave-functionals in QFTs we assume can be represented by Euclidean path-integrals in the lower-half-planes. As a result, the reduced density matrix elements on subsystem AA can be written in terms of the Euclidean path-integrals with branch-cuts on AA:

α|ρA|β=Φ(A)=βΦ(A+)=α[𝒟Φ]eIE(g,Φ)\langle\alpha|\rho_{A}|\beta\rangle=\int_{\Phi\left(A^{-}\right)=\beta}^{\Phi\left(A^{+}\right)=\alpha}[\mathcal{D}\Phi]e^{-I_{E}\left(g,\Phi\right)} (10)

where Φ\Phi denotes the quantum fields collectively. Now we perform an infinitesimal shape deformation of the entangling region AA~A\rightarrow\tilde{A}. For such a deformation, we can always find a corresponding coordinates transformation :xμx~μ(x)=xμ+ζμ(x)\mathcal{F}:x^{\mu}\to\tilde{x}^{\mu}(x)=x^{\mu}+\zeta^{\mu}(x) on the whole Euclidean space-time such that (A~)=A\mathcal{F}(\tilde{A})=A (see Fig 2).

Refer to caption
Figure 2: Relating shape deformation to metric deformation via a corresponding diffeomorphism.

Via diffeomorphism equivalence we can then trade the shape-deformation for a metric deformation gμνg~μνg_{\mu\nu}\to\tilde{g}_{\mu\nu} and write the reduced density matrix elements as:

α|ρA~g|β=α~|ρAg~|β~=Φ(A)=β~Φ(A+)=α~[𝒟Φ]eIE(g~,Φ)\langle\alpha|\rho_{\tilde{A}}^{g}|\beta\rangle=\langle\tilde{\alpha}|\rho_{A}^{\tilde{g}}|\tilde{\beta}\rangle=\int_{\Phi\left(A^{-}\right)=\tilde{\beta}}^{\Phi\left(A^{+}\right)=\tilde{\alpha}}[\mathcal{D}\Phi]e^{-I_{E}\left(\tilde{g},\Phi\right)} (11)

where:

α~=α,β~=β,g~μν=μx~ανx~βgαβ\tilde{\alpha}=\alpha\circ\mathcal{F},\;\;\tilde{\beta}=\beta\circ\mathcal{F},\;\;\tilde{g}_{\mu\nu}=\nabla_{\mu}\tilde{x}^{\alpha}\nabla_{\nu}\tilde{x}^{\beta}g_{\alpha\beta} (12)

For deformations ζμ(x)\zeta^{\mu}(x) that are small, we can extract the change in the reduced density matrix to leading order in ζ\zeta. They are simply given by linear response of the metric deformation, i.e. proportional to the stress tensor:

δρA\displaystyle\delta\rho_{A} =\displaystyle= UρA~UρA=12δgμνρAT^μν+𝒪(ζ2)\displaystyle U^{\dagger}\circ\rho_{\tilde{A}}\circ U-\rho_{A}=\frac{1}{2}\int\delta g^{\mu\nu}\rho_{A}\;\hat{T}_{\mu\nu}+\mathcal{O}\left(\zeta^{2}\right)
δgμν\displaystyle\delta g^{\mu\nu} =\displaystyle= 2μζν,U|α~=|α\displaystyle 2\nabla^{\mu}\zeta^{\nu},\;\;U|\tilde{\alpha}\rangle=|\alpha\rangle (13)

2.2 Shape response from the entanglement first law

In order to compute the change in entanglement entropy, recall that under any change of the reduced density matrix: ρ~AρA+λδρA,λ1\tilde{\rho}_{A}\to\rho_{A}+\lambda\;\delta\rho_{A},\;\lambda\ll 1, the first order change is dictated by the so-called the entanglement first law:

δSA=λtrδρAHA+𝒪(λ2),HA=lnρA\delta S_{A}=\lambda\;\text{tr}\delta\rho_{A}H_{A}+\mathcal{O}(\lambda^{2}),\;\;H_{A}=-\ln{\rho_{A}} (14)

This is a consequence of the positivity constraint for the relative entropy between ρA\rho_{A} and ρ~A\tilde{\rho}_{A}. Combining the entanglement first law with the shape perturbation theory (13), one can write the linear response of entanglement entropy under shape deformation as:

δSA\displaystyle\delta S_{A} =\displaystyle= SA~SA=tr(UρA~U)ln(UlnρA~U)+trρAlnρA\displaystyle S_{\tilde{A}}-S_{A}=-\text{tr}\left(U^{\dagger}\circ\rho_{\tilde{A}}\circ U\right)\ln{\left(U^{\dagger}\circ\ln{\rho_{\tilde{A}}}\circ U\right)}+\text{tr}\rho_{A}\ln{\rho_{A}} (15)
=\displaystyle= trδρAHA+𝒪(ζ2)=𝑑xgμζν(x)tr{ρATμν(x)HA}+𝒪(ζ2)\displaystyle\text{tr}\;\delta\rho_{A}\;H_{A}+\mathcal{O}\left(\zeta^{2}\right)=\int dx\sqrt{g}\,\ \nabla^{\mu}\zeta^{\nu}(x)\;\text{tr}\left\{\rho_{A}T_{\mu\nu}(x)H_{A}\right\}+\mathcal{O}(\zeta^{2})
=\displaystyle= 𝑑xgμζν(x)Tμν(x)HA+𝒪(ζ2)\displaystyle\int dx\sqrt{g}\,\ \nabla^{\mu}\zeta^{\nu}(x)\;\left\langle T_{\mu\nu}(x)H_{A}\right\rangle+\mathcal{O}(\zeta^{2})

We see that in order to apply (15), we need to have the knowledge of modular Hamiltonians, and compute its correlation functions with the stress tensor.

2.3 Warming up: shape dependence of single-interval entanglement entropy in 2d CFT

As a warm-up exercise for subsequent computations, we revisit the shape dependence of single-interval entanglement entropy in 2d CFT using the entanglement first law. The result is of course known, and the computation for single-sphere in arbitrary dimensional CFTs has been done, e.g. in Mezei2015Feb ; Mezei2015Feb02 . The purpose is to recognize patterns and subtleties that may appear in the computations for mutual information.

Refer to caption
Figure 3: The UV-cut surface at the A\partial{A} of the single interval (R,R)(-R,R) in 2D CFT. For the deformation vector field ζ(x)\zeta(x), we only show its restriction on A\partial A which is relevant for our calculations.

We label the single-interval subregion AA by:

A{RxR},τ=0A\equiv\{-R\leq x\leq R\},\quad\tau=0 (16)

From Casini2011May , the modular Hamiltonian for AA is given by:

HA=2πAR2x22RT00(x)𝑑xH_{A}=2\pi\int_{A}\frac{R^{2}-x^{2}}{2R}T_{00}(x)dx (17)

Now we deform the single-interval subregion AA by moving its right end point RR+ζx(R)R\rightarrow R+\zeta^{x}(R). This can be realized by a deformation vector field ζ(x)\zeta(x) supported about the end point RR. The linear response of SAS_{A} under this deformation can be computed by (15):

δSA\displaystyle\delta S_{A} =d2xμζν(x)Tμν(x)HA\displaystyle=\int d^{2}x\,\ \partial^{\mu}\zeta^{\nu}(x)\left\langle T_{\mu\nu}(x)H_{A}\right\rangle (18)
=d2xμ(ζν(x)Tμν(x)HA)d2xζν(x)μTμν(x)HA\displaystyle=\int d^{2}x\,\ \partial^{\mu}\bigg{(}\zeta^{\nu}(x)\left\langle T_{\mu\nu}(x)H_{A}\right\rangle\bigg{)}-\int d^{2}x\,\ \zeta^{\nu}(x)\left\langle\partial^{\mu}T_{\mu\nu}(x)H_{A}\right\rangle
=d2xμ(ζν(x)Tμν(x)HA)=𝑑x¯hnμζν(x¯)Tμν(x¯)HA\displaystyle=\int d^{2}x\,\ \partial^{\mu}\bigg{(}\zeta^{\nu}(x)\left\langle T_{\mu\nu}(x)H_{A}\right\rangle\bigg{)}=\int_{\mathcal{E}}d\bar{x}\sqrt{h}\,\ n^{\mu}\zeta^{\nu}(\bar{x})\left\langle T_{\mu\nu}(\bar{x})H_{A}\right\rangle

In the second line, we perform an integration by parts, the second term vanishes by Ward-identity up to possible contact terms. In this case, we can use the local nature of modular flow by HAH_{A} and infer that:

μTμν(x)HA\displaystyle\left\langle\partial^{\mu}T_{\mu\nu}(x)H_{A}\right\rangle =eiHAsμTμν(x)eiHAsHAμTμν(xs)HA\displaystyle=\left\langle e^{iH_{A}s}\partial^{\mu}T_{\mu\nu}(x)e^{-iH_{A}s}H_{A}\right\rangle\propto\left\langle\partial^{\mu}T_{\mu\nu}(x_{s})H_{A}\right\rangle (19)

where xsx_{s} is the modular flow trajectory of (17). Therefore one can always apply appropriate modular flow to move μTμν\partial^{\mu}T_{\mu\nu} away from the support of HAH_{A}, and conclude that there will be no contact term contribution. We can therefore drop this term. As a total derivative, the first term is reduced to a boundary integral by Stokes theorem. The boundary is imposed as a UV regulator for the calculation of entanglement entropy. In this case, the boundary \mathcal{E} is a codimension one tube of radius ϵ\epsilon enclosing the entangling surface A\partial A, i.e. =A×S1\mathcal{E}=\partial A\times S^{1}. For our case, the entangling surface A\partial A is just the right end point. Now we proceed by plugging in the explicit form of HAH_{A} into (18), then the boundary integral becomes the contour integral in z=x+iτz=x+i\tau:

δSA\displaystyle\delta S_{A} =iz=Rdz2πζz(z)T(z)HA+h.c.\displaystyle=i\oint_{z=R}\frac{dz}{2\pi}\;\zeta^{z}(z)\left\langle T(z)H_{A}\right\rangle+h.c. (20)
=iζz(R)z=Rdz2πRR𝑑w(R2w22R)T(z)T(w)+h.c.\displaystyle=i\zeta^{z}(R)\oint_{z=R}\frac{dz}{2\pi}\int_{-R}^{R}dw\;\left(\frac{R^{2}-w^{2}}{2R}\right)\langle T(z)T(w)\rangle+h.c.
=iζz(R)z=Rdz2πRR𝑑w(R2w22R)c/2(zw)4+h.c.\displaystyle=i\zeta^{z}(R)\oint_{z=R}\frac{dz}{2\pi}\int_{-R}^{R}dw\;\left(\frac{R^{2}-w^{2}}{2R}\right)\frac{c/2}{(z-w)^{4}}+h.c.
=iζz(R)z=Rdz2π[cR23(Rz)2(R+z)2]+h.c.\displaystyle=i\zeta^{z}(R)\oint_{z=R}\frac{dz}{2\pi}\;\left[\frac{cR^{2}}{3(R-z)^{2}(R+z)^{2}}\right]+h.c.
=c6R(ζz(R)+ζz¯(R))=c6Rζx(R)\displaystyle=\frac{c}{6R}\left(\zeta^{z}(R)+\zeta^{\bar{z}}(R)\right)=\frac{c}{6R}\zeta^{x}(R)

which agrees with the well-known result:

SA=c3log(2Rϵ)S_{A}=\frac{c}{3}\log\left(\frac{2R}{\epsilon}\right) (21)

In the second line of (20) , we have replaced ζz(z)\zeta^{z}(z) by its average ζz(R)\zeta^{z}(R) because we are only interested in the mode that translates the tube \mathcal{E}, instead of changing the geometry of the tube which the nonzero modes of ζ\zeta effectuate.

We summarize the take-away message of (20) as follows. In the context of the entanglement first law, the linear response of entanglement entropy is proportional to a contour integral of stress tensor correlator with modular Hamiltonian:

δSAzA𝑑zT(z)HA\delta S_{A}\propto\oint_{z\sim\partial A}{dz}\;\langle T(z)H_{A}\rangle (22)

This only picks up the residue of simple pole in T(z)HA\langle T(z)H_{A}\rangle, schematically of the form:

T(z)HA=+ReszA+\langle T(z)H_{A}\rangle=\cdots+\frac{Res}{z-\partial A}+\cdots (23)

In cases where the modular Hamiltonian can be written as integrals of local operators like (17), such poles emerge only after completing all the integrals in the modular Hamiltonian. For example, switching the order of 𝑑w\int{dw} and z=A𝑑z\oint_{z=\partial A}{dz} in (20) gives zero instead. This point is important later when we study shape dependence of mutual information using the entanglement first law.

3 Mutual information and modular Hamiltonian in the OPE limit

Our goal is to study the shape dependence of mutual information perturbatively about two spheres in the CFT vacuum. Similar to the computation of δSA\delta S_{A} in Sec (2.3), this can be done via the entanglement first law. The pre-requisite for such an approach is explicit knowledge of the unperturbed result for both the mutual information and its modular Hamiltonians. However, mutual information in general is highly theory dependent, even for the case of our interest, i.e. two spheres in the vacuum. Explicit results are only accessible in the so-called OPE limit, in which the distance between two spheres are much larger than their radii. As a result, in this paper we shall also focus on the shape dependence of mutual information in the same (or conformally equivalent) limit (see Fig 1). In this section, we review the results in Faulkner2016Aug ; Faulkner2021Aug calculations of for the OPE limit of mutual information and the corresponding modular Hamiltonian using the replica technique.

We consider two disjoint spherical subregions AA and BB with radius RA,RBR_{A},R_{B} separated by a distance LL (see Fig 1). The mutual information IA,BI_{A,B} can be obtained by the replica trick as: IA,Blimn1In(A,B)I_{A,B}\equiv\lim_{n\rightarrow 1}I_{n}(A,B), where:

In(A,B)Sn(A)+Sn(B)Sn(AB)=11nlog(trρAntrρBntrρABn)I_{n}(A,B)\equiv S_{n}(A)+S_{n}(B)-S_{n}(AB)=\frac{1}{1-n}\log\left(\frac{\operatorname{tr}\rho_{A}^{n}\operatorname{tr}\rho_{B}^{n}}{\operatorname{tr}\rho_{AB}^{n}}\right) (24)

is the Rényi mutual information. For QFTs defined on Euclidean spacetime \mathcal{M}, we can rewrite these traces as partition functions on appropriately constructed conifolds, i.e.

In(A,B)=1n1log(Z(𝒞nAB)Zn()Z(𝒞nA)Z(𝒞nB))I_{n}(A,B)=\frac{1}{n-1}\log\left(\frac{Z\left(\mathcal{C}_{n}^{A\cup B}\right)Z^{n}(\mathcal{M})}{Z\left(\mathcal{C}_{n}^{A}\right)Z\left(\mathcal{C}_{n}^{B}\right)}\right) (25)

where 𝒞nX\mathcal{C}_{n}^{X} is constructed from nn-copies of \mathcal{M}, glued across subregion XX. Equivalently, we can write the partition functions on conifold as expectation values of twist operators in the corresponding orbifold theory: CFTn/nCFT^{\otimes n}/\mathds{Z}_{n}, defined on n\mathcal{M}^{n}:

Z(𝒞nA)=ZnΣnAn,Z(𝒞nB)=ZnΣnBn\displaystyle Z\left(\mathcal{C}_{n}^{A}\right)=Z^{n}\left\langle\Sigma_{n}^{A}\right\rangle_{\mathcal{M}^{n}},\quad Z\left(\mathcal{C}_{n}^{B}\right)=Z^{n}\left\langle\Sigma_{n}^{B}\right\rangle_{\mathcal{M}^{n}} (26)
Z(𝒞nAB)=ZnΣnAΣnBn.\displaystyle Z\left(\mathcal{C}_{n}^{A\cup B}\right)=Z^{n}\left\langle\Sigma_{n}^{A}\Sigma_{n}^{B}\right\rangle_{\mathcal{M}^{n}}.

Analogous to ordinary correlation functions, the twist operators in ΣnAΣnBn\left\langle\Sigma_{n}^{A}\Sigma_{n}^{B}\right\rangle_{\mathcal{M}^{n}} also admit operator product expansions into local operators in CFTn/nCFT^{\otimes n}/\mathds{Z}_{n} :

ΣnA,B\displaystyle\Sigma_{n}^{A,B} =Z(𝒞nA,B)Zn(1+{kj}C{kj}A,Bj=0n1Φkj(j)(rA,B))\displaystyle=\frac{Z\left(\mathcal{C}_{n}^{A,B}\right)}{Z^{n}}\bigg{(}1+\sum_{\left\{k_{j}\right\}}C_{\left\{k_{j}\right\}}^{A,B}\prod_{j=0}^{n-1}\Phi_{k_{j}}^{(j)}\left(r_{A,B}\right)\bigg{)} (27)

i.e. expanding ΣnA,B\Sigma_{n}^{A,B} into a complete set of local operators at rA,Br_{A,B}, where in (27) we have written the leading identity operator separately. Assuming a 2\mathds{Z}_{2} symmetry for the underlying CFT, the coefficients for single copy operators vanish. As a result, the first few orders in the OPE take the form :

ΣnA,B=ΣnA,Bn(1+12(2RA,B)2Δjkn1cjk𝒪(j)𝒪(k)+)\Sigma_{n}^{A,B}=\left\langle\Sigma_{n}^{A,B}\right\rangle_{\mathcal{M}^{n}}\bigg{(}1+\frac{1}{2}(2R_{A,B})^{2\Delta}\sum_{j\neq k}^{n-1}c_{j-k}\mathcal{O}^{(j)}\mathcal{O}^{(k)}+\cdots\bigg{)} (28)

where 𝒪\mathcal{O} is a primary operator with the lowest scaling dimension Δ\Delta, and it enters the OPE by appearing twice in different replica jkj\neq k. We call such form of OPE the bi-local channel. Due to replica symmetry, the OPE coefficients only depend on jkj-k. The subsequent \cdots denotes contributions from other operators in the bi-local channel or from multi-local channels. For future convenience, we have rescaled the OPE coefficients with a factor (2RA,B)2Δ(2R_{A,B})^{-2\Delta} relative to (27), i.e. cjk=(2RA,B)2ΔCjkc_{j-k}=(2R_{A,B})^{-2\Delta}C_{j-k}. In this paper, we focus on the leading-order contribution to IA,BI_{A,B}. This corresponds to 𝒪(j)𝒪(k)\mathcal{O}^{(j)}\mathcal{O}^{(k)} composed by the lowest-dimension primary operator 𝒪\mathcal{O}. For simplicity, we begin by assuming 𝒪\mathcal{O} as a scalar, and will discuss the cases for spinning operators later.

Analogous to the usual OPE, we can extract the coefficients cjkc_{j-k} by inserting corresponding operators at rr\to\infty and inverting (27):

cjk\displaystyle c_{j-k} =\displaystyle= limrΣnA,B𝒪(j)(r)𝒪k(r)G1(r)G1(r)\displaystyle\lim_{r\to\infty}\left\langle\Sigma^{A,B}_{n}\mathcal{O}^{(j)}(r)\mathcal{O}^{k}(r)\right\rangle G^{-1}(r)G^{-1}(r) (29)
=\displaystyle= limrΩjΔΩkΔG1(r)G1(r)𝒪(Zj)𝒪(Zk)𝕊(2πn)×d1\displaystyle\lim_{r\to\infty}\Omega_{j}^{\Delta}\;\Omega_{k}^{\Delta}\;G^{-1}(r)\;G^{-1}(r)\;\left\langle\mathcal{O}\left(Z_{j}\right)\mathcal{O}\left(Z_{k}\right)\right\rangle_{\mathds{S}_{(2\pi n)}\times\mathds{H}_{d-1}}
=\displaystyle= 𝒪(τj,0)𝒪(τk,0)𝕊(2πn)×d1\displaystyle\left\langle\mathcal{O}(\tau_{j},0)\mathcal{O}(\tau_{k},0)\right\rangle_{\mathds{S}_{(2\pi n)}\times\mathds{H}_{d-1}}
=\displaystyle= Gn(τiτj),τi=(2i+1)π\displaystyle G_{n}(\tau_{i}-\tau_{j}),\;\;\tau_{i}=(2i+1)\pi

The correlation function in the presence of ΣnA,B\Sigma^{A,B}_{n} has been written in terms of thermal correlator GnG_{n} in hyperbolic space d1\mathds{H}_{d-1} at inverse temperature T1=2πnT^{-1}=2\pi n. This is achieved via the coordinate transformation into hyperbolic space coordinates Z=(τ,Zi)𝕊(2πn)×d1Z=(\tau,Z^{i})\in\mathds{S}_{(2\pi n)}\times\mathds{H}_{d-1}:

Z0\displaystyle Z^{0} =\displaystyle= τ=cos1(R2r2(R2r2)2+4R2r02)\displaystyle\tau=\cos^{-1}\left(\frac{R^{2}-r^{2}}{\sqrt{(R^{2}-r^{2})^{2}+4R^{2}r_{0}^{2}}}\right)
Zi\displaystyle Z^{i} =\displaystyle= 2Rri(R2r2)2+4R2r02,i1\displaystyle\frac{2Rr_{i}}{\sqrt{(R^{2}-r^{2})^{2}+4R^{2}r_{0}^{2}}},\;\;i\geq 1 (30)

From this form, one can easily perform the analytic continuation in nn for the replica trick and obtain the contribution to the mutual information:

IA,B=limn111n(jkcjk2)(2RA2RBL2)2Δ\displaystyle I_{A,B}=\lim_{n\to 1}\frac{1}{1-n}\bigg{(}\sum_{j\neq k}c_{j-k}^{2}\bigg{)}\left(\frac{2R_{A}2R_{B}}{L^{2}}\right)^{2\Delta} (31)
=\displaystyle= (2RA2RBL2)2Δ𝑑sG1(is+π)24cosh2(s/2)\displaystyle\left(\frac{2R_{A}2R_{B}}{L^{2}}\right)^{2\Delta}\int^{\infty}_{-\infty}ds\frac{G_{1}(-is+\pi)^{2}}{4\cosh^{2}\left(s/2\right)}
=\displaystyle= (2RA2RBL2)2Δπ42Δ+1Γ(2Δ+1)Γ(2Δ+3/2).\displaystyle\left(\frac{2R_{A}2R_{B}}{L^{2}}\right)^{2\Delta}\frac{\sqrt{\pi}}{4^{2\Delta+1}}\frac{\Gamma(2\Delta+1)}{\Gamma(2\Delta+3/2)}.

The prefactor is just the conformal ratio ρ\rho in (9) in the OPE limit. More details of the computations can be found in Cardy2013 ; Faulkner2016Aug ; Faulkner2021Aug .

For the application of the entanglement first law, we also need the explicit form of the modular Hamiltonians in the combination ΔHA,BHA,BHAHB\Delta H_{A,B}\equiv H_{A,B}-H_{A}-H_{B}. In the OPE limit, it admits an expansion similar to that of the mutual information. In this case the OPE is only taken for one of the twist operators ΣnB\Sigma_{n}^{B}. After analytic continuation, the local operators from OPE are modular flown by HAH_{A}. The details can be found in Faulkner2021Aug . In the end, one obtains the OPE for ΔHA,B\Delta H_{A,B} in the bi-local channel of primary scalar operator 𝒪\mathcal{O} as:

ΔHA,B\displaystyle\Delta H_{A,B} =\displaystyle= (2RB)2Δ𝑑sk1(s)𝒪A(is)𝒪B\displaystyle-{(2R_{B})^{2\Delta}}\int_{-\infty}^{\infty}{ds}\,\ k_{1}(s)\;\mathcal{O}_{A}(-is)\;\mathcal{O}_{B}
+\displaystyle+ i(2RB)2Δ2π𝑑sj𝑑skk2(sj,sk)𝒪A(iskisj)𝒪A(isj)+δHBB\displaystyle\frac{i(2R_{B})^{2\Delta}}{2\pi}\int_{-\infty}^{\infty}{ds_{j}ds_{k}}\,\ k_{2}(s_{j},s_{k})\;\mathcal{O}_{A}(-is_{k}-is_{j})\;\mathcal{O}_{A}(-is_{j})+\delta H_{BB}
k1(s)\displaystyle k_{1}(s) =\displaystyle= 14cosh2(s/2)c1(is+π)\displaystyle\frac{1}{4\cosh^{2}(s/2)}c_{1}(-is+\pi)
k2(sj,sk)\displaystyle k_{2}(s_{j},s_{k}) =\displaystyle= 14cosh2(sj/2)(1esk+iϵ1+1esk+sj+1)c1(isk+ϵ)\displaystyle\frac{1}{4\cosh^{2}(s_{j}/2)}\left(\frac{1}{e^{s_{k}+i\epsilon}-1}+\frac{1}{e^{s_{k}+s_{j}}+1}\right)c_{1}(-is_{k}+\epsilon) (32)

where c1(is)c_{1}(is) is the analytic continuation of cnc_{n} at n=1n=1:

c1(is)=(4sinh2(s/2))Δc_{1}(is)=\left(-4\sinh^{2}(s/2)\right)^{-\Delta} (33)
Refer to caption
Figure 4: The locations for 𝒪A(is)\mathcal{O}_{A}(-is) and 𝒪B\mathcal{O}_{B}. In the limit of large separation, the modular trajectory of 𝒪A(is)\mathcal{O}_{A}(-is) almost passes through the center of AA.

The operator 𝒪A(is)\mathcal{O}_{A}(-is) is obtained from 𝒪B=𝒪(rB)\mathcal{O}_{B}=\mathcal{O}(r_{B}) by first applying a modular reflection in HAH_{A}, followed with a modular flow by real ss (see Fig 4):

𝒪A(is)=ρAis1/2𝒪BρAis+1/2.\mathcal{O}_{A}(-is)=\rho_{A}^{is-1/2}\;\mathcal{O}_{B}\;\rho_{A}^{-is+1/2}.

As a check, one can compute the expectation value of the modular Hamiltonian (3) in the vacuum state:

ΔHA,B=ΔHA,B(1)+ΔHA,B(2)\displaystyle\langle\Delta H_{A,B}\rangle=\langle\Delta H_{A,B}^{(1)}\rangle+\langle\Delta H_{A,B}^{(2)}\rangle (34)

where ΔHA,B(1)\langle\Delta H_{A,B}^{(1)}\rangle and ΔHA,B(1)\langle\Delta H_{A,B}^{(1)}\rangle represents the contribution from the single integral and double integral of Eq (3) respectively. It can be found that:

ΔHA,B(1)=2ΔHA,B(2)=2ρ2Δπ42Δ+1Γ(2Δ+1)Γ(2Δ+3/2)=2IA,B.\displaystyle\langle\Delta H_{A,B}^{(1)}\rangle=-2\langle\Delta H_{A,B}^{(2)}\rangle=-2\rho^{2\Delta}\frac{\sqrt{\pi}}{4^{2\Delta+1}}\frac{\Gamma(2\Delta+1)}{\Gamma(2\Delta+3/2)}=-2I_{A,B}. (35)

which once combined in (34) agrees with the result for mutual information (31).

4 Shape dependence of mutual information in the OPE limit

In this section, we give the main calculations of this paper. The task is to compute the linear response of mutual information IA,BI_{A,B} between two spheres AA and BB under shape deformation on one of the spheres AA~A\to\tilde{A}. The spheres have radius RAR_{A} and RBR_{B} whose centers are separated by a distance LL. We focus on a conformally equivalent version of the OPE limit LRARBL\sim R_{A}\gg R_{B} (see Figure 5).

Refer to caption
Figure 5: The setup of spherical frame coordinates YY. The center of the sphere AA has been set to the origin. While that of sphere BB is located at Y1=L,Yi=0Y_{1}=L,Y_{i}=0. The angular coordinate α\alpha on A\partial A is defined as the angle relative to the Y1Y_{1} axis.

More specifically, the deformed sphere A~\tilde{A} can be parametrized by an angular-dependent modification of the radius: R(Ω)=RA+ζ(Ω)R(\Omega)=R_{A}+\zeta(\Omega). In this parametrization, we can work with the ansatz in which the deformation vector field ζμ\zeta^{\mu} that enters the entanglement first law (15) only has a radial component in polar coordinates about the center of A:ζμδrμA:\zeta^{\mu}\propto\delta^{\mu}_{r}.

Now, let us calculate the shape deformation of mutual information by using the ingredients that we have prepared in previous. Similar to the warm-up exercise in 2.3, application of the entanglement first law yields that:

δIA,B\displaystyle\delta I_{A,B} =ddYμζν(Y)Tμν(Y)ΔHA,B\displaystyle=\int_{\mathcal{M}}d^{d}Y\,\ \partial^{\mu}\zeta^{\nu}(Y)\left\langle T_{\mu\nu}(Y)\Delta H_{A,B}\right\rangle (36)
=ddYμ(ζν(Y)Tμν(Y)ΔHA,B)ddYζν(Y)μTμν(Y)ΔHA,B\displaystyle=\int_{\mathcal{M}}d^{d}Y\,\ \partial^{\mu}\bigg{(}\zeta^{\nu}(Y)\left\langle T_{\mu\nu}(Y)\Delta H_{A,B}\right\rangle\bigg{)}-\int_{\mathcal{M}}d^{d}Y\,\ \zeta^{\nu}(Y)\left\langle\partial^{\mu}T_{\mu\nu}(Y)\Delta H_{A,B}\right\rangle
=ϵ𝑑θdd2Ωh(Ω)nμζνTμν(θ,Ω)ΔHA,B,\displaystyle=\int_{\partial\mathcal{M}}\epsilon\;{d\theta}\;{d}^{d-2}\Omega\sqrt{h(\Omega)}\,\ n^{\mu}\zeta^{\nu}\langle T_{\mu\nu}(\theta,\Omega)\Delta H_{A,B}\rangle,

where S1×Sd2\partial\mathcal{M}\sim S^{1}\times S^{d-2} is the UV cut-off surface in the form of codimension one tube of radius ϵ\epsilon enclosing the spherical boundary A\partial A, and (θ,Ωd2)(\theta,\Omega_{d-2}) are coordinates on \partial\mathcal{M}. Similar to (18), here we can also drop the contact term in the second line. The reason is that the OPE of ΔHA,B\Delta H_{A,B} consists of bi-local operators whose location we can avoid by restricting the support of ζμ\zeta^{\mu} in the bulk.

From Eq. (3), we see that the OPE limit of modular Hamiltonian ΔHA,B\Delta H_{A,B} is written as the integral of local operators which are flowed by the modular Hamiltonian of spherical subregion AA. In terms of actual calculations, this is not convenient because the expressions for the modular trajectories are very complicated. We can simplify the calculations by exploiting the conformal symmetry. More specifically, we observe that (36) involves integrals of CFT correlators of the form

Tμν𝒪(is1)𝒪(is2)=TμνρAis𝒪(YB)ρAis1+is2𝒪(YB)ρAis2\langle T_{\mu\nu}\mathcal{O}(-is_{1})\mathcal{O}(-is_{2})\rangle=\langle T_{\mu\nu}\rho_{A}^{is}\mathcal{O}(Y_{B})\rho_{A}^{-is_{1}+is_{2}}\mathcal{O}(Y_{B})\rho_{A}^{-is_{2}}\rangle (37)

where YBY_{B} is the center of a small sphere BB. For CFTs, there is a conformal transformation 𝒰\mathcal{U} that maps from AA to the Rindler half plane. The modular evolution operator can be related to Rindler boost KK by:

ρAis=𝒰1eisK𝒰\rho_{A}^{is}=\mathcal{U}^{-1}e^{isK}\mathcal{U} (38)

Plugging (38) into (37), and using the invariance of the conformal vacuum under 𝒰\mathcal{U}, we can rewrite the correlator in the Rindler frame:

Tμν𝒪(is1)𝒪(is2)\displaystyle\langle T_{\mu\nu}\mathcal{O}(-is_{1})\mathcal{O}(-is_{2})\rangle =Tμν𝒰1eis1K𝒰𝒪𝒰1eis1K+is2K𝒰𝒪𝒰1eis2K𝒰\displaystyle=\langle T_{\mu\nu}\mathcal{U}^{-1}e^{is_{1}K}\mathcal{U}\mathcal{O}\mathcal{U}^{-1}e^{-is_{1}K+is_{2}K}\mathcal{U}\mathcal{O}\mathcal{U}^{-1}e^{-is_{2}K}\mathcal{U}\rangle (39)
=𝒰Tμν𝒰1eis1K𝒰𝒪𝒰1eis1K+is2K𝒰𝒪𝒰1eis2K\displaystyle=\langle\mathcal{U}T_{\mu\nu}\mathcal{U}^{-1}e^{is_{1}K}\mathcal{U}\mathcal{O}\mathcal{U}^{-1}e^{-is_{1}K+is_{2}K}\mathcal{U}\mathcal{O}\mathcal{U}^{-1}e^{-is_{2}K}\rangle
=T~μνeis1K𝒪~eis1K+is2K𝒪~eis2K\displaystyle=\langle\tilde{T}_{\mu\nu}e^{is_{1}K}\tilde{\mathcal{O}}e^{-is_{1}K+is_{2}K}\tilde{\mathcal{O}}e^{-is_{2}K}\rangle
=T~μν𝒪~(is1)𝒪~(is2)\displaystyle=\langle\tilde{T}_{\mu\nu}\tilde{\mathcal{O}}(-is_{1})\tilde{\mathcal{O}}(-is_{2})\rangle

where T~μν=𝒰1Tμν𝒰\tilde{T}_{\mu\nu}=\mathcal{U}^{-1}\;T_{\mu\nu}\;\mathcal{U}, and 𝒪~(is)=Λ(YB)ΔeiKs𝒪(𝒰(YB))eiKs\tilde{\mathcal{O}}(-is)=\Lambda(Y_{B})^{-\Delta}\;e^{iKs}\mathcal{O}\Big{(}\mathcal{U}(Y_{B})\Big{)}e^{-iKs} are boosted operators in Rindler frame rescaled by the conformal factor Λ(YB)\Lambda(Y_{B}) of 𝒰\mathcal{U}.

4.1 Mapping spherical frame to Rindler frame

The conformal transformation 𝒰\mathcal{U} from the spherical frame coordinates YY to the Rindler frame coordinate XX is given explicitly by the following form:

CIXμ\displaystyle C_{I}X^{\mu} =Y~μ+(Y~Y~)Cμ1+2(CY~)+(CC)(Y~Y~),\displaystyle=\frac{\tilde{Y}^{\mu}+\big{(}\tilde{Y}\cdot\tilde{Y}\big{)}C^{\mu}}{1+2\big{(}C\cdot\tilde{Y}\big{)}+\big{(}C\cdot C\big{)}\big{(}\tilde{Y}\cdot\tilde{Y}\big{)}}, (40)
Y~μ\displaystyle\tilde{Y}^{\mu} =YμRAδ1μ,Cμ=δ1μ/2RA\displaystyle=Y^{\mu}-R_{A}\delta^{\mu}_{1},\quad C^{\mu}={\delta_{1}^{\mu}}/{2R_{A}}
CI\displaystyle C_{I} =2RA(L2RA2RB2)(L+RA+RB)(L+RARB)\displaystyle=\frac{2R_{A}\big{(}L^{2}-R_{A}^{2}-R_{B}^{2}\big{)}}{\big{(}L+R_{A}+R_{B}\big{)}\big{(}L+R_{A}-R_{B}\big{)}}

The map is arranged such that under 𝒰\mathcal{U} the spherical subregion AA becomes the left half space region x1<0x^{1}<0 and the spherical subregion BB becomes a new sphere with new center at the canonical position XB=(0,1,0,,0)X_{B}=(0,1,0,\cdots,0) and new radius is given by (see Fig 6):

Refer to caption
Figure 6: Rindler frame: subregion AA is the half space X1<0X_{1}<0; subregion BB is a sphere whose center is located at XB=(0,1,0,,0)X_{B}=(0,1,0,...,0) with small radius RI1R_{I}\ll 1.
RI=2RARB(L2RA2RB2)=ρ2\displaystyle R_{I}=\frac{2R_{A}R_{B}}{\left(L^{2}-R_{A}^{2}-R_{B}^{2}\right)}=\frac{\rho}{2} (41)

The corresponding conformal factor is given by:

Λ(Y)=CI(1+2(CY~)+C2Y~2)\Lambda(Y)=C_{I}\left(1+2(C\cdot\tilde{Y})+C^{2}\tilde{Y}^{2}\right)

To utilize (39), we need to keep track of how the UV cut-off surface transform across 𝒰\mathcal{U}. To proceed, we parametrize the cut-off tube in spherical frame YμY^{\mu} (see Fig 5) as:

Y0=ϵsinθ,Y1=(RA+ϵcosθ)cosα\displaystyle Y^{0}=\epsilon\sin\theta,\qquad Y_{1}=\Big{(}R_{A}+\epsilon\cos\theta\Big{)}\cos\alpha (42)
Yi=(RA+ϵcosθ)(sinα)ei,i2ei2=1,\displaystyle Y_{i}=\Big{(}R_{A}+\epsilon\cos\theta\Big{)}(\sin\alpha)e_{i},\qquad\sum_{i\geq 2}e_{i}^{2}=1,

where ϵ\epsilon is the UV cut-off scale. Under the transformation of (40), the UV cut-off manifold S1×Sd2S^{1}\times S^{d-2} is mapped to a tube S1×d2S^{1}\times\mathds{R}^{d-2} by the following relation:

CIX0=ϵ2sinθ1+cosα+𝒪(ϵ2)\displaystyle C_{I}X^{0}=\epsilon\frac{2\sin\theta}{1+\cos\alpha}+\mathcal{O}\left(\epsilon^{2}\right) (43)
CIX1=ϵ2cosθ1+cosα+𝒪(ϵ2)\displaystyle C_{I}X_{1}=\epsilon\frac{2\cos\theta}{1+\cos\alpha}+\mathcal{O}\left(\epsilon^{2}\right)
CIXi=2RAsinα1+cosαei+𝒪(ϵ2),\displaystyle C_{I}X_{i}=\frac{2R_{A}\sin\alpha}{1+\cos\alpha}e_{i}+\mathcal{O}\left(\epsilon^{2}\right),
ei2=1\displaystyle\sum e_{i}^{2}=1

The geometry of the mapped tube in Rindler frame is changed such that it has XiX_{i}-dependent radius (see Fig 7):

Refer to caption
Figure 7: The UV cut-off tube in two conformally equivalent frames. Left: the UV cut-off tube of constant radius in the original spherical frame. Right: the mapped UV cut-off tube of “trumpet” shape in the Rindler frame.
ϵ~(Xi)=ϵ(1+CI24RA2i2Xi2)CI1\tilde{\epsilon}(X_{i})=\epsilon\Big{(}1+\frac{C_{I}^{2}}{4R_{A}^{2}}\sum_{i\geq 2}X_{i}^{2}\Big{)}{C_{I}}^{-1} (44)

where (0,0,Xi)(0,0,X_{i}) is a corresponding point on the Rindler plane d2\mathds{R}^{d-2} .

Now, we can plug the explicit form of 𝒰\mathcal{U} into (39), the correlator Tμν𝒪(is1)𝒪(is2)\langle T_{\mu\nu}\mathcal{O}(-is_{1})\mathcal{O}(-is_{2})\rangle can be explicit written as boosted correlators in Rindler frame:

nμζν(θ,Ω)Tμν(θ,Ω)𝒪(s1)𝒪(s2)\displaystyle n^{\mu}\zeta^{\nu}(\theta,\Omega)\langle T_{\mu\nu}(\theta,\Omega)\mathcal{O}(s_{1})\mathcal{O}(s_{2})\rangle
=Λ(YB)2ΔΛ(θ,Ω)(d2)(dXαdYμdXβdYν)|θ,Ωnμζν(θ,Ω)Tαβ(θ,Ω)𝒪(X[s1])𝒪(X[s2])\displaystyle={\Lambda(Y_{B})}^{-2\Delta}{\Lambda(\theta,\Omega)}^{-(d-2)}\left(\frac{dX^{\alpha}}{dY^{\mu}}\frac{dX^{\beta}}{dY^{\nu}}\right)\Big{|}_{\theta,\Omega}n^{\mu}\zeta^{\nu}(\theta,\Omega)\left\langle T_{\alpha\beta}({\theta,\Omega})\mathcal{O}(X[s_{1}])\mathcal{O}(X[s_{2}])\right\rangle

where

X[s]±=e±s,Xi=0X[s]^{\pm}=e^{\pm s},\quad X^{i}=0 (45)

are the boosted coordinates, and the conformal factors are given by:

Λ(YB)=CI(L+RA)24RA2\displaystyle\Lambda(Y_{B})=\frac{C_{I}\left(L+R_{A}\right)^{2}}{4R_{A}^{2}} (46)
Λ(θ,Ω)=CI(1+CI2i2Xi2(θ,Ω)4RA2)1\displaystyle\Lambda(\theta,\Omega)=C_{I}\left(1+C_{I}^{2}\frac{\sum_{i\geq 2}X_{i}^{2}(\theta,\Omega)}{4R_{A}^{2}}\right)^{-1}

The factor Λ(YB)2Δ\Lambda(Y_{B})^{-2\Delta} will combine with the (2RB)2Δ(2R_{B})^{2\Delta} in the OPE coefficient to form the conformal ratio ρ2Δ\rho^{2\Delta} in the limit RB0R_{B}\to 0; the factor Λ(θ,Ω)\Lambda(\theta,\Omega) reflects the geometry new cut-off tube encoded in (44). Gather all pieces together, we can rewrite δIA,B\delta I_{A,B} as follows:

δIA,B\displaystyle\delta I_{A,B} =ϵ~𝑑θdd2Xin~μξνTμν(θ,Ω)ΔH~A,B\displaystyle=\int\tilde{\epsilon}{d}\theta\int{d}^{d-2}X_{i}\,\ \tilde{n}^{\mu}\xi^{\nu}\left\langle T_{\mu\nu}(\theta,\Omega)\Delta\tilde{H}_{A,B}\right\rangle (47)

where n~μ=eiθzeiθz¯\tilde{n}^{\mu}=-e^{i\theta}\partial_{z}-e^{-i\theta}\partial_{\bar{z}}, z=X1+iX0,z¯=X1iX0z=X^{1}+iX^{0},\,\ \bar{z}=X^{1}-iX^{0} is just the unit normal vector for new UV cut-off tube, and

ξμ(X)=ξ(X)(z+z¯),ξ(X)=(1+CI24RA2i2Xi2(θ,Ω))CI1ζ(θ,Ω)\xi^{\mu}(X)=\xi(X)\left(\partial_{z}+\partial_{\bar{z}}\right),\quad\xi(X\small)=\Big{(}1+\frac{C_{I}^{2}}{4R_{A}^{2}}\sum_{i\geq 2}X_{i}^{2}(\theta,\Omega)\Big{)}{C_{I}}^{-1}{\zeta}(\theta,\Omega) (48)

is the “effective” deformation vector field that appears in the Rindler frame. Furthermore, ΔH~A,B\Delta\tilde{H}_{A,B} is now in terms of boosted operators in Rindler frame:

ΔH~A,B=\displaystyle\Delta\tilde{H}_{A,B}= ρ2Δ𝑑sk1(s)𝒪(X[s])𝒪(X[0])\displaystyle-\rho^{2\Delta}\int_{-\infty}^{\infty}ds\;k_{1}(s)\;\mathcal{O}\left(-X[s]\right)\mathcal{O}\left(X[0]\right) (49)
+iρ2Δ2π𝑑sj𝑑skk2(sk,sj)𝒪(X[sj+sk])𝒪(X[sj])\displaystyle+\frac{i\rho^{2\Delta}}{2\pi}\int_{-\infty}^{\infty}ds_{j}\;ds_{k}\;k_{2}\left(s_{k},s_{j}\right)\;\mathcal{O}\left(-X[s_{j}+s_{k}]\right)\mathcal{O}\left(-X[s_{j}]\right)

Therefore, all computations can be carried out in the Rindler frame, we simply need to keep track of the mapping between conformally related deformation vector fields (48).

4.2 Shape dependence in Rindler frame

Now, we compute the shape dependence of mutual information (47) by working in the Rindler frame, where the modular flow becomes simply the boost about the half plane X1<0X_{1}<0. Similar to the warm-up exercise, we can write the integral on the cut-off tube in terms of contour integrals in z=X1+iX0,z¯=X1iX0z=X^{1}+iX^{0},\,\ \bar{z}=X^{1}-iX^{0} :

δIA,B\displaystyle\delta I_{A,B} =idd2Xiξ(Xi)|z|=ϵ~𝑑z(Tzz(z,z¯,Xi)+Tzz¯(z,z¯,Xi))ΔH~A,B+h.c.\displaystyle=i\int{d}^{d-2}X^{i}\,\ \xi(X^{i})\oint_{|z|=\tilde{\epsilon}}{d}z\,\ \left\langle\bigg{(}T_{zz}(z,\bar{z},X^{i})+T_{z\bar{z}}(z,\bar{z},X^{i})\bigg{)}\Delta\tilde{H}_{A,B}\right\rangle+h.c. (50)

Notice that in writing (50), we have taken the deformation vector field ξ(Xi)\xi(X^{i}) outside the contour integral 𝑑z\oint{dz} by the zero-mode substitution:

ξ(z,z¯,Xi)ξ(Xi)dz2πizξ(z,z¯,Xi)\xi(z,\bar{z},X^{i})\to\xi(X^{i})\equiv\oint\frac{dz}{2\pi iz}\xi(z,\bar{z},X^{i})

The reason is the same as in (20), i.e. we only want the mode that moves the cut-off tube instead of deforming it.

The universal data regarding the linear response δIA,B\delta I_{A,B} is encoded in the following “susceptibility” of shape deformation:

δIA,Bδξ(Xi)=(i)|z|=ϵ~𝑑z(Tzz(z,z¯,Xi)+Tzz¯(z,z¯,Xi))ΔH~A,B+h.c.\displaystyle\frac{\delta I_{A,B}}{\delta\xi(X^{i})}=(i)\oint_{|z|=\tilde{\epsilon}}{d}z\,\ \left\langle\bigg{(}T_{zz}(z,\bar{z},X^{i})+T_{z\bar{z}}(z,\bar{z},X^{i})\bigg{)}\Delta\tilde{H}_{A,B}\right\rangle+h.c. (51)

which simply picks up the residue of simple poles in TzzΔH~A,B\langle T_{zz}\Delta\tilde{H}_{A,B}\rangle, etc. Unlike in the two-dimensional case of the warm-up exercise, strictly speaking the integrands are neither holomorphic nor anti-holomorphic functions of zz and z¯\bar{z}. However, in the limit ϵ0\epsilon\to 0 the contributions from the integrands can be approximated by corresponding holomorphic or anti-holomorphic functions, see Appendix A.

The basic building blocks for the integrands are three points functions T𝒪𝒪\langle T\mathcal{O}\mathcal{O}\rangle. In dd dimensional CFTs, they are completely fixed by conformal symmetry to have the form Simmons-Duffin2016 :

𝒪(X1)𝒪(X2)Tμν(X3)=C12THμν(X1,X2,X3)|X12|2Δd+2|X13|d2|X23|d2,\left\langle\mathcal{O}\left(X_{1}\right)\mathcal{O}\left(X_{2}\right)T^{\mu\nu}\left(X_{3}\right)\right\rangle=C_{12T}\frac{H^{\mu\nu}\left(X_{1},X_{2},X_{3}\right)}{\left|X_{12}\right|^{2\Delta-d+2}\left|X_{13}\right|^{d-2}\left|X_{23}\right|^{d-2}}, (52)
Hμν\displaystyle H^{\mu\nu} =VμVν1dVαVαδμν,Vμ=X13μX132X23μX232\displaystyle=V^{\mu}V^{\nu}-\frac{1}{d}V_{\alpha}V^{\alpha}\delta^{\mu\nu},\quad V^{\mu}=\frac{X_{13}^{\mu}}{X_{13}^{2}}-\frac{X_{23}^{\mu}}{X_{23}^{2}} (53)
C12T\displaystyle C_{12T} =dΔd11SdC12,Sd=2πd/2Γ(d/2).\displaystyle=-\frac{d\Delta}{d-1}\frac{1}{S_{d}}C_{12},\quad S_{d}=\frac{2\pi^{d/2}}{\Gamma(d/2)}.

where C12C_{12} is the normalization constant of the two-point function 𝒪1(x1)𝒪2(x2)\left\langle\mathcal{O}_{1}\left(x_{1}\right)\mathcal{O}_{2}\left(x_{2}\right)\right\rangle which we set to 11. The details of doing integrals are complicated so we package them in Appendix A where interested readers can refer to. We will instead make a few general comments and observations below.

Due to the general form of the modular Hamiltonian ΔH~A,B\Delta\tilde{H}_{A,B}, there are two corresponding integrals for δIA,B\delta I_{A,B}, schematically of the form:

δIA,B(1)\displaystyle\delta{I}_{A,B}^{(1)}\sim 𝑑z𝑑sk1(s)T𝒪(X[s])𝒪(X[0]),\displaystyle\oint{dz}\;\int{ds}\;k_{1}(s)\left\langle T\mathcal{O}(X[s])\mathcal{O}(X[0])\right\rangle, (54)
δIA,B(2)\displaystyle\delta{I}_{A,B}^{(2)}\sim 𝑑z𝑑sk𝑑sjk2(sk,sj)T𝒪(X[sj+sk])𝒪(X[sj])\displaystyle\oint{dz}\;\int{ds_{k}}\;\int{ds_{j}}\;k_{2}(s_{k},s_{j})\left\langle T\mathcal{O}(X[s_{j}+s_{k}])\mathcal{O}(X[s_{j}])\right\rangle

Analogous to HAH_{A} for the single-interval case (17), ΔH~A,B\Delta\tilde{H}_{A,B} is also written as modular integral (or double integrals) of local operators. Therefore, according to section 2.3 we should expect the corresponding simple poles in zz to emerge only after completing all the integrals in modular parameters. It turns out the single integral 𝑑s\int{ds} in δIA,B(1)\delta I_{A,B}^{(1)} does not produce simple poles of zz, thus its contribution is zero. The response comes entirely from the double integral 𝑑sk𝑑sj\int{ds_{k}}\int{ds_{j}} in δIA,B(2)\delta I_{A,B}^{(2)}. However, it is slightly surprising that one does not need to complete both integrals for simple poles in zz to emerge; only one of the integrals 𝑑sj\int ds_{j} is sufficient. The response is then obtained by integrating the sks_{k}-dependent residue over the remaining modular parameter. In the end, we obtain the following result:

δIA,B=NΔ,ddd2Xiξ(Xi)(1+D)d+1\displaystyle\delta I_{A,B}=N_{\Delta,d}\int{d}^{d-2}X^{i}\,\ \xi(X^{i})(1+D)^{-d+1} (55)

where D=i=2d1(Xi)2D=\sum_{i=2}^{d-1}(X^{i})^{2} and NΔ,d=Δρ2Δπ42Δ+1Γ(2Δ+1)Γ(2Δ+3/2)2d1Γ(d12)πd/21/2N_{\Delta,d}=\Delta\rho^{2\Delta}\frac{\sqrt{\pi}}{4^{2\Delta+1}}\frac{\Gamma(2\Delta+1)}{\Gamma(2\Delta+3/2)}\frac{2^{d-1}\Gamma\left(\frac{d-1}{2}\right)}{\pi^{d/2-1/2}}.

4.3 Mapping back to spherical frame

Now that we have the results in Rindler frame, it is simple to translate them back in the spherical frame and obtain the general results on spheres. Writing (55) in terms of the original deformation field ζ\zeta by (47), we have that:

δIA,B=NΔ,ddd2Xiζ(X)CI1(1+CI24RA2i2Xi2)(1+i2Xi2)d+1\displaystyle\delta I_{A,B}=N_{\Delta,d}\int{d}^{d-2}X^{i}\,\ {\zeta}(X)C_{I}^{-1}\bigg{(}1+\frac{C_{I}^{2}}{4R_{A}^{2}}\sum_{i\geq 2}X_{i}^{2}\bigg{)}\bigg{(}1+\sum_{i\geq 2}X_{i}^{2}\bigg{)}^{-d+1} (56)

The last step is to rewrite (56) in terms of spherical coordinates (α,ei)(\alpha,e_{i}) on sphere AA. The integral measure now takes the form:

dd2Xi\displaystyle\int{d}^{d-2}X_{i} =𝑑Ωd2YRAd2CI(d2)(1+CI24RA2i2Xi2)d2,\displaystyle=\int{d}{{\Omega}_{d-2}^{Y}}\,\ R_{A}^{d-2}C_{I}^{-(d-2)}\Big{(}1+\frac{C_{I}^{2}}{4R_{A}^{2}}\sum_{i\geq 2}X_{i}^{2}\Big{)}^{d-2}, (57)

From (43), we have that:

i2Xi2=4RA2CI21cosα1+cosα\sum_{i\geq 2}X_{i}^{2}=\frac{4R_{A}^{2}}{C_{I}^{2}}\frac{1-\cos\alpha}{1+\cos\alpha}

Then combining everything, we finally obtain the result for the linear response of mutual information between spheres, for the OPE contribution from scalar primary operator of scaling dimension Δ\Delta, it takes the form :

δIA,BΔ,=0=N~d(IA,BΔ,=0lnρ)𝑑Ωd2(L2RA2L2+RA22LRAcosα)d1ζ(Ω)\displaystyle\delta I_{A,B}^{\Delta,\ell=0}=\tilde{N}_{d}\;\bigg{(}\frac{\partial I_{A,B}^{\Delta,\ell=0}}{\partial\ln\rho}\bigg{)}\int{d}{{\Omega}_{d-2}}\,\ \left(\frac{L^{2}-R_{A}^{2}}{L^{2}+R_{A}^{2}-2LR_{A}\cos\alpha}\right)^{d-1}{\zeta}(\Omega) (58)

where N~d=(RASd1)1\tilde{N}_{d}=\left(R_{A}S_{d-1}\right)^{-1} and we have made the identification:

2Δρ2Δπ42Δ+1Γ(2Δ+1)Γ(2Δ+3/2)=(IA,BΔ,=0lnρ)2\Delta\rho^{2\Delta}\frac{\sqrt{\pi}}{4^{2\Delta+1}}\frac{\Gamma(2\Delta+1)}{\Gamma(2\Delta+3/2)}=\left(\frac{\partial I_{A,B}^{\Delta,\ell=0}}{\partial\ln\rho}\right) (59)

One can plug in (58) special deformations of the form:

ζd(Ω)=ζ0,ζt(Ω)=ζ0cosα{\zeta}^{d}(\Omega)=\zeta_{0},\;\;{\zeta}^{t}(\Omega)=\zeta_{0}\cos\alpha (60)

which generates dilation RARA+ζ0R_{A}\to R_{A}+\zeta_{0} and translation LLζ0L\to L-\zeta_{0} of the sphere AA. We can compute for these deformations:

δζdIA,BΔ,=0=(ζ0RAL2+RA2L2RA2)(IA,BΔ,=0lnρ)\displaystyle\delta_{\zeta^{d}}I_{A,B}^{\Delta,\ell=0}=\left(\frac{\zeta_{0}}{R_{A}}\frac{L^{2}+R_{A}^{2}}{L^{2}-R_{A}^{2}}\right)\left(\frac{\partial I_{A,B}^{\Delta,\ell=0}}{\partial\ln\rho}\right) (61)
δζtIA,BΔ,=0=(2ζ0LL2RA2)(IA,BΔ,=0lnρ)\displaystyle\delta_{\zeta^{t}}I_{A,B}^{\Delta,\ell=0}=\left(\frac{2\zeta_{0}L}{L^{2}-R_{A}^{2}}\right)\left(\frac{\partial I_{A,B}^{\Delta,\ell=0}}{\partial\ln\rho}\right)

which agrees with the result obtained by varying RAR_{A} and LL directly in (31) .

Eq (58) is the main result of our paper. The most important feature is a form factor that depends only on the space-time dimension dd, and is otherwise quite universal, e.g. independent of the scaling dimension Δ\Delta. In addition, the “susceptibility” of shape deformation (refered to as simply susceptibility from now on) is only proportional to powers of the distance between the center of small ball BB and deformation point Ω\Omega on AA:

δIA,BΔ,=0h(Ω)δζ(Ω)|YBYΩ|2(d1)\frac{\delta I^{\Delta,\ell=0}_{A,B}}{\sqrt{h(\Omega)}\delta\zeta(\Omega)}\propto|Y_{B}-Y_{\Omega}|^{-2(d-1)} (62)

Therefore, the shape deformation only induces a response that corresponds to the “zero-mode” in a sense that we will discuss later in section 5.

At this point, these observations only pertains to a particular type of contribution, i.e. those from scalar primary operators. In principle, the susceptibility could still depend on the spin \ell of the primary operator forming the bi-local OPE channel. It is therefore worth studying the effect of including spins to the computation of shape response. Conducting an exhaustive investigation of all spinning contributions is however beyond the scope of our goal, instead we study the case with spin-1 primary operator, and observe if the corresponding susceptibility is modified.

4.4 Including spins: case study for spin-1 contributions

We now study the OPE contribution to mutual information IA,BI_{A,B} from the bi-local operator consisting of two spin-1 primary operators with conformal dimension Δ\Delta:

ΣnB1×ΣnB12(2RB)2Δjkcjkμν𝒪μ(j)𝒪ν(k)\left\langle\Sigma^{B}_{n}\right\rangle^{-1}\times\Sigma^{B}_{n}\rightarrow\frac{1}{2}(2R_{B})^{2\Delta}\sum_{j\neq k}c^{\mu\nu}_{j-k}\mathcal{O}_{\mu}^{(j)}\mathcal{O}_{\nu}^{(k)} (63)

The goal is to compute the linear response of this contribution under shape deformation. Similar to the scalar case, the OPE coefficients cjkμνc^{\mu\nu}_{j-k} for spinning primary operator is related to the thermal correlator of spinning operators in the hyperbolic space d1\mathds{H}_{d-1} at inverse temperature T1=2πnT^{-1}=2\pi n using (3):

cjkμν\displaystyle c^{\mu\nu}_{j-k} =\displaystyle= limrΩiΔ1ΩjΔ1(dZjdr)αρ(dZkdr)βσG1(r)ρμG1(r)σν𝒪α(Zj)𝒪β(Zk)𝕊(2πn)×d1\displaystyle\lim_{r\to\infty}\Omega_{i}^{\Delta-1}\Omega_{j}^{\Delta-1}\left(\frac{dZ_{j}}{dr}\right)^{\rho}_{\alpha}\left(\frac{dZ_{k}}{dr}\right)^{\sigma}_{\beta}G^{-1}(r)^{\mu}_{\rho}G^{-1}(r)^{\nu}_{\sigma}\left\langle\mathcal{O}^{\alpha}(Z_{j})\mathcal{O}^{\beta}(Z_{k})\right\rangle_{\mathds{S}_{(2\pi n)}\times\mathds{H}_{d-1}} (64)
=\displaystyle= Gnμν(τjτk),τi=(2i+1)π\displaystyle G^{\mu\nu}_{n}(\tau_{j}-\tau_{k}),\;\;\tau_{i}=(2i+1)\pi

A systematic treatment of replica trick for OPE into spinning primaries as well as their contributions to the mutual information can be found in Casini2021Sep , see also ChenBin2017 ; Long2016 . In order to compute its shape response using the entanglement first law, we need to know its OPE contribution to the modular Hamiltonian. Although the derivation in Faulkner2021Aug was applied explicitly to the scalar case (3), the properties used in the derivation (KMS condition, translation invariance in replica indices) are valid for cjkμνc^{\mu\nu}_{j-k} as well. As a result, one can simply substitute c1(is)c1μν(is)c_{1}(is)\to c^{\mu\nu}_{1}(is) and thus obtain the spin-1 contribution to the modular Hamiltonian:

ΔHA,BΔ,=1\displaystyle\Delta H^{\Delta,\ell=1}_{A,B} =\displaystyle= ρ2Δ𝑑sk1μν(s)ρAis𝒪μ(XB)ρAis𝒪ν(XB)\displaystyle-{\rho^{2\Delta}}\int_{-\infty}^{\infty}{ds}\,\ k^{\mu\nu}_{1}(s)\rho_{A}^{is}\mathcal{O}_{\mu}(-X_{B})\rho_{A}^{-is}\mathcal{O}_{\nu}(X_{B})
+\displaystyle+ iρ2Δ2π𝑑sj𝑑skk2μν(sj,sk)ρAi(sj+sk)𝒪μ(XB)ρAisk𝒪ν(XB)ρAisj+δHII\displaystyle\frac{i\rho^{2\Delta}}{2\pi}\int_{-\infty}^{\infty}{ds_{j}ds_{k}}\,\ k^{\mu\nu}_{2}(s_{j},s_{k})\rho_{A}^{-i(s_{j}+s_{k})}\mathcal{O}_{\mu}(-X_{B})\;\rho_{A}^{is_{k}}\mathcal{O}_{\nu}(-X_{B})\;\rho_{A}^{is_{j}}+\delta H_{II}
k1μν(s)\displaystyle k^{\mu\nu}_{1}(s) =\displaystyle= 14cosh2(s/2)c1μν(is+π)\displaystyle\frac{1}{4\cosh^{2}(s/2)}c^{\mu\nu}_{1}(-is+\pi)
k2μν(sj,sk)\displaystyle k^{\mu\nu}_{2}(s_{j},s_{k}) =\displaystyle= 14cosh2(sj/2)(1esk+iϵ1+1esk+sj+1)c1μν(isk+ϵ)\displaystyle\frac{1}{4\cosh^{2}(s_{j}/2)}\left(\frac{1}{e^{s_{k}+i\epsilon}-1}+\frac{1}{e^{s_{k}+s_{j}}+1}\right)c^{\mu\nu}_{1}(-is_{k}+\epsilon) (65)

The coefficient c1μν(is)c_{1}^{\mu\nu}(is) is the analytically continued correlator in hyperbolic space at n=1n=1, which is conformally related to flat-space correlator:

c1μν(isis)=𝒪μ(is,0)𝒪ν(is,0)𝕊2π×d1\displaystyle c^{\mu\nu}_{1}\left(is-is^{\prime}\right)=\left\langle\mathcal{O}^{\mu}(is,0)\mathcal{O}^{\nu}(is^{\prime},0)\right\rangle_{\mathds{S}_{2\pi}\times\mathds{H}_{d-1}} (66)
=\displaystyle= limr(r(is)Z)αμ(r(is)Z)βν𝒪α(r(is))𝒪β(r(is))d\displaystyle\lim_{r\to\infty}\left(\frac{\partial r(is)}{\partial Z}\right)^{\mu}_{\alpha}\left(\frac{\partial r(is^{\prime})}{\partial Z}\right)^{\nu}_{\beta}\left\langle\mathcal{O}^{\alpha}(r(is))\mathcal{O}^{\beta}(r(is^{\prime}))\right\rangle_{\mathds{R}^{d}}

The trajectories r(is)r(is) are the images of hyperbolic time translation ττ+is\tau\to\tau+is in d\mathds{R}^{d} starting at r(0)=rr(0)=r. They are nothing but the modular-flow trajectories:

r(s)=R((r2R2)sinh(s),2Rri)2(r2sinh2(s2)R2cosh2(s2))r(s)=\frac{R\left((r^{2}-R^{2})\sinh{(s)},-2Rr_{i}\right)}{2\left(r^{2}\sinh^{2}{\left(\frac{s}{2}\right)}-R^{2}\cosh^{2}{\left(\frac{s}{2}\right)}\right)} (67)

We need to take the rr\to\infty limit to extract the OPE coefficient, in which case the trajectories simplify to:

limrr(s)=(Rcoth(s2),0)\lim_{r\to\infty}r(s)=\left(R\coth{\left(\frac{s}{2}\right)},\vec{0}\right) (68)

We can then obtain that:

c1μν(is)=c1(is)(ημν2n^μn^ν)c^{\mu\nu}_{1}(is)=c_{1}(is)\left(\eta^{\mu\nu}-2\hat{n}^{\mu}\hat{n}^{\nu}\right) (69)

where c1c_{1} is the scalar coefficient and n^μ=nμ/(nn)1/2\hat{n}^{\mu}=n^{\mu}/(n\cdot n)^{1/2} is the normalized time-like vector pointing from the center of the sphere to the tip of its null cone. In our case ημ=δ0μ\eta^{\mu}=\delta^{\mu}_{0} but the formula works more generally for boosted spheres. The dependence on n^μ\hat{n}^{\mu} in the OPE coefficients encodes the orientation of the twist operator ΣnI\Sigma^{I}_{n}, as in the usual OPE expansion. Plugging (69) into (4.4), the contribution of spin-1 primary to mutual information can be found to be dd times that of scalar primary:

IA,BΔ,=1=dρ2Δπ42Δ+1Γ(2Δ+1)Γ(2Δ+3/2)I^{\Delta,\ell=1}_{A,B}=d\rho^{2\Delta}\frac{\sqrt{\pi}}{4^{2\Delta+1}}\frac{\Gamma(2\Delta+1)}{\Gamma(2\Delta+3/2)} (70)

This agrees with results in Casini2021Sep ; ChenBin2017 ; Long2016 .

Now one can proceed and compute the linear response of (4.4) to shape deformation using the entanglement first law. The strategy is identical to the scalar computation, and one still finds that only the double integral term in (4.4) produces a response. We shall not repeat the details and instead zoom into the key computation, which is the following integral:

δIA,B(2)𝑑z𝑑sj𝑑skk2μν(sj,sk)Tzz(z,z¯;Xi)ρAi(sj+sk)𝒪μ(XB)ρAisk𝒪ν(XB)ρAisj\delta I^{(2)}_{A,B}\propto\oint dz\int^{\infty}_{-\infty}ds_{j}ds_{k}k^{\mu\nu}_{2}(s_{j},s_{k})\left\langle T_{zz}(z,\bar{z};X^{i})\rho_{A}^{-i(s_{j}+s_{k})}\mathcal{O}_{\mu}(-X_{B})\;\rho_{A}^{is_{k}}\mathcal{O}_{\nu}(-X_{B})\;\rho_{A}^{is_{j}}\right\rangle

For spinning operators, the modular flow is also accompanied by a tensor factor:

ρAis𝒪μ(X)ρAis=ΩsΔ1R(s)μα𝒪α(X[s])\displaystyle\rho_{A}^{-is}\mathcal{O}_{\mu}(X)\rho_{A}^{is}=\Omega_{s}^{\Delta-1}R(s)^{\alpha}_{\mu}\;\mathcal{O}_{\alpha}\left(X[s]\right) (71)

In the Rindler frame where we evaluate the integral, it takes the form of the boost:

R(s)μα=esδμ+δ+α+esδμδα+δμα|α,μ2R(s)^{\alpha}_{\mu}=e^{s}\delta^{+}_{\mu}\delta^{\alpha}_{+}+e^{-s}\delta^{-}_{\mu}\delta^{\alpha}_{-}+\delta^{\alpha}_{\mu}|_{\alpha,\mu\geq 2} (72)

For spinning operators, one also needs the tensor structure QzzμνQ_{zz\mu\nu} appearing in the relevant three-point function:

𝒪μ(x1)𝒪ν(x2)Tzz(x3)=Qμνzz(x12,x13,x23)|x12|2Δd|x13|d2|x23|d2\left\langle\mathcal{O}_{\mu}(x_{1})\mathcal{O}_{\nu}(x_{2})T_{zz}(x_{3})\right\rangle=\frac{Q_{\mu\nu zz}(x_{12},x_{13},x_{23})}{|x_{12}|^{2\Delta-d}|x_{13}|^{d-2}|x_{23}|^{d-2}} (73)

Conformal symmetry restricts the form of the tensor structures QQ to be a linear combination of 4 basic building blocks, which can be most compactly encoded using the embedding space coordinates PP and the corresponding auxiliary null vector TT Costa2011 :

Q\displaystyle Q =\displaystyle= a1V1,23V2,31V3,122+a2V3,122H12\displaystyle a_{1}V_{1,23}V_{2,31}V_{3,12}^{2}+a_{2}V_{3,12}^{2}H_{12} (74)
+\displaystyle+ a3(V1,23H23+V2,31H13)V3,12+a4H13H23\displaystyle a_{3}\left(V_{1,23}H_{23}+V_{2,31}H_{13}\right)V_{3,12}+a_{4}H_{13}H_{23}

The building blocks Vi,jkV_{i,jk} and HijH_{ij} are defined by:

Vi,jk\displaystyle V_{i,jk} =\displaystyle= (TiPj)(PiPk)(TiPk)(PiPj)(PjPk)\displaystyle\frac{(T_{i}\cdot P_{j})(P_{i}\cdot P_{k})-(T_{i}\cdot P_{k})(P_{i}\cdot P_{j})}{(P_{j}\cdot P_{k})}
Hij\displaystyle H_{ij} =\displaystyle= 2[(TiTj)(PiPj)(TiPj)(TjPi)]\displaystyle-2\left[(T_{i}\cdot T_{j})(P_{i}\cdot P_{j})-(T_{i}\cdot P_{j})(T_{j}\cdot P_{i})\right] (75)

To read out the components QμνzzQ_{\mu\nu zz}, one first re-write the contractions in terms of physical coordinates xx and the corresponding auxiliary vector tt using:

TiTjtitj,PiPj12xij2,PiTjtjxij\displaystyle T_{i}\cdot T_{j}\to t_{i}\cdot t_{j},\;\;P_{i}\cdot P_{j}\to-\frac{1}{2}x_{ij}^{2},\;\;P_{i}\cdot T_{j}\to t_{j}\cdot x_{ij} (76)

and then extract the coefficient of t1μt2νt3zt3zt^{\mu}_{1}t^{\nu}_{2}t^{z}_{3}t^{z}_{3}. The computation proceeds more tediously but is otherwise identical to the scalar case, so we spare the details and simply write the result in Rindler frame:

δIA,BΔ,=1\displaystyle\delta I^{\Delta,\ell=1}_{A,B} =\displaystyle= NΔ,d=1dd2X(1+i2Xi2)(d1)ζ(X)\displaystyle N_{\Delta,d}^{\ell=1}\int{d^{d-2}X}\;\left(1+\sum_{i\geq 2}X_{i}^{2}\right)^{-(d-1)}\zeta(X)
NΔ,d=1\displaystyle N_{\Delta,d}^{\ell=1} =\displaystyle= 2d1Sd1(IA,BΔ,=1lnρ)(d1)Sdd2Δ(a1a2d2a3+2a4)\displaystyle\frac{2^{d-1}}{S_{d-1}}\left(\frac{\partial I_{A,B}^{\Delta,\ell=1}}{\partial\ln\rho}\right)\frac{(d-1)S_{d}}{d^{2}\Delta}\left(a_{1}-a_{2}d-2a_{3}+2a_{4}\right) (77)

In the spherical frame, we would then have that:

δIA,B=1=(21dRA)NΔ,d=1𝑑Ωd2(L2RA2L2+RA22LRAcosα)d1ζ(Ω)\delta I^{\ell=1}_{A,B}=\left(\frac{2^{1-d}}{R_{A}}\right)N_{\Delta,d}^{\ell=1}\int{d\Omega_{d-2}}\;\left(\frac{L^{2}-R_{A}^{2}}{L^{2}+R_{A}^{2}-2LR_{A}\cos\alpha}\right)^{d-1}{\zeta}(\Omega) (78)

We can see that the susceptibility contains the same form factor as the scalar contribution in (58). The coefficients ai,i=1,,4a_{i},i=1,...,4 are generally theory-dependent. In our case, they are further constrained by the conservation of TμνT_{\mu\nu} and its Ward identity. This will reduce the number of independent coefficients, but there will still be free parameters that are theory-dependent. In particular, consistency with (70) for special deformations (60) then requires that:

a1a2d2a3+2a4=d2Δ(d1)Sda_{1}-a_{2}d-2a_{3}+2a_{4}=\frac{d^{2}\Delta}{(d-1)S_{d}} (79)

so that:

δIA,BΔ,=1=N~d(IA,BΔ,=1lnρ)𝑑Ωd2(L2RA2L2+RA22LRAcosα)d1ζ(Ω)\delta I^{\Delta,\ell=1}_{A,B}=\tilde{N}_{d}\;\left(\frac{\partial I_{A,B}^{\Delta,\ell=1}}{\partial\ln\rho}\right)\int{d}{{\Omega}_{d-2}}\,\ \left(\frac{L^{2}-R_{A}^{2}}{L^{2}+R_{A}^{2}-2LR_{A}\cos\alpha}\right)^{d-1}{\zeta}(\Omega) (80)

For the special case of 𝒪μ\mathcal{O}_{\mu} being a conserved current with conformal dimension Δ=d1\Delta=d-1, the coefficients can be more explicitly written in Osborn1994 ; Hofman2016 ; Dymarsky2019 as:

a1\displaystyle a_{1} =\displaystyle= (23d4d3γ)C,a2=(12d4d2γ)C\displaystyle(2-3d-4d^{3}\gamma)C,\;\;a_{2}=(1-2d-4d^{2}\gamma)C
a3\displaystyle a_{3} =\displaystyle= 2d(1+4γ)C,a4=2(1d24γd)C,C=d(d2)2(d1)2Sd\displaystyle-2d(1+4\gamma)C,\;\;a_{4}=2\left(\frac{1}{d-2}-4\gamma d\right)C,\;\;C=\frac{d(d-2)}{2(d-1)^{2}S_{d}} (81)

where we have normalized the current two-point function. In this case one can check that the free parameter γ\gamma cancels out and (79) is indeed satisfied. In principle, (79) should be derivable in CFTs from the conformal ward identity directly. In our context it is imposed for the shape dependence of mutual information, which is obtained from Tμν𝒪α𝒪β\langle T_{\mu\nu}\mathcal{O}_{\alpha}\mathcal{O}_{\beta}\rangle, to be consistent with the change in mutual information between spheres δIδ𝒪α𝒪β\delta I\sim\delta\langle\mathcal{O}_{\alpha}\mathcal{O}_{\beta}\rangle under conformal transformations (60) – this is essentially the conformal ward identity. Since the scalar and spin-1 contributions (31, 70) to the mutual information between spheres are themselves highly universal – they only depend on the conformal ratio, this also explains partially the universality we observe in their shape-dependences.

We conclude that for the bi-local OPE contribution from spin =1\ell=1 primary operators, the susceptibility under shape deformation reveals the same form factor as the scalar result (58). Therefore, it is worth conjecturing that the susceptibility is of the universal form:

δIA,BΔ,=N~d(IA,BΔ,lnρ𝑑Ωd2)(L2RA2L2+RA22LRAcosα)d1ζ(Ω)\delta I^{\Delta,\ell}_{A,B}=\tilde{N}_{d}\;\left(\frac{\partial I^{\Delta,\ell}_{A,B}}{\partial\ln\rho}\int{d}{{\Omega}_{d-2}}\right)\,\ \left(\frac{L^{2}-R_{A}^{2}}{L^{2}+R_{A}^{2}-2LR_{A}\cos\alpha}\right)^{d-1}{\zeta}(\Omega) (82)

for all bi-local OPE contributions from primary operators, i.e. independent of both their conformal dimension Δ\Delta and spin \ell.

5 Extremization of mutual information in the OPE limit

In this section, we try to unpack some implications of (58). In the presence of symmetries, the response theory can be diagonalized into eigenmodes of the symmetry generators. For example, in Mezei2015Feb ; Mezei2015Feb02 the linear response under shape deformation of entanglement entropy of spherical sub-regions in CFT vacuum has been found to only invoke the zero mode of spherical harmonics (i.e. the average) multiplied by a cut-off dependence – an indication for the extremization of its universal part on spheres:

δSδ(d2)Sd2𝑑Ωd2ζ(Ω)\delta S\propto\delta^{-(d-2)}\int_{S_{d-2}}d\Omega_{d-2}\zeta(\Omega) (83)

In this case the spherical harmonics should emerge as the natural candidate for mode decomposition of the response, this is due to the rotational symmetry of the set up. To clarify more, the role of rotational symmetry has two aspects: firstly, it is the symmetry of the underlying CFT, and in particular of its vacuum; secondly, the entangling surface (thus the defect twist operator Σn\Sigma_{n} inserted when computing via the replica trick) is kept invariant under the action of symmetry transformation. To be more precise, the vacuum is symmetric under rotations about any point; the spherical defect is kept invariant by rotations about a particular point as the center – the “little group” of Σn\Sigma_{n}.

Now with regard to mutual information between two spherical sub-regions AA and BB, the “little group” is generically empty in the presence of two spherical defects ΣnA\Sigma^{A}_{n} and ΣnB\Sigma^{B}_{n} for vacuum in generic relativistic QFTs. However for CFTs we still have a non-trivial little group, i.e. symmetries of the CFT vacuum that keep both sub-regions invariant. The crucial ingredient is a special conformal transformation 𝒦\mathcal{K} that maps two spheres to be positioned concentrically (see Fig 8).

Refer to caption
Figure 8: Conformally equivalent configuration of two concentric spheres.

More concretely, we denote the centers of two spheres by YAY_{A} and YBY_{B} respectively, separated by distance LL. The special conformal transformation 𝒦\mathcal{K} consists of two inversions across carefully chosen origins. The first inversion 1\mathcal{I}_{1} across the origin Y0Y_{0} needs to make BB cover AA co-centrically. We can parametrize Y0Y_{0} by λ\lambda:

Y0=λYA+(1λ)YB,  0<λ<1\displaystyle Y_{0}=\lambda Y_{A}+(1-\lambda)Y_{B},\;\;0<\lambda<1 (84)

The inversion 1\mathcal{I}_{1} maps the two spheres to be centered at:

Y~A\displaystyle\tilde{Y}_{A} =\displaystyle= Y0(1λ)(YBYA)(1λ)2L2RA2\displaystyle Y_{0}-\frac{(1-\lambda)\left(Y_{B}-Y_{A}\right)}{(1-\lambda)^{2}L^{2}-R_{A}^{2}}
Y~B\displaystyle\tilde{Y}_{B} =\displaystyle= Y0λ(YBYA)RB2λ2L2\displaystyle Y_{0}-\frac{\lambda\left(Y_{B}-Y_{A}\right)}{R_{B}^{2}-\lambda^{2}L^{2}} (85)

Solving for the co-centricity condition Y~A=Y~B\tilde{Y}_{A}=\tilde{Y}_{B} gives that:

λ=12(1+rB2rA2(1+rB2rA2)24rB2),rA,B=RA,BL\lambda=\frac{1}{2}\left(1+r_{B}^{2}-r_{A}^{2}-\sqrt{\left(1+r_{B}^{2}-r_{A}^{2}\right)^{2}-4r_{B}^{2}}\right),\;\;r_{A,B}=\frac{R_{A,B}}{L} (86)

so that the common center Y~c=Y~A,B\tilde{Y}_{c}=\tilde{Y}_{A,B} is located at:

Y~c\displaystyle\tilde{Y}_{c} =\displaystyle= λ~YA+(1λ~)YB\displaystyle\tilde{\lambda}Y_{A}+(1-\tilde{\lambda})Y_{B}
λ~\displaystyle\tilde{\lambda} =\displaystyle= λ+1L2((1+rB2rA2)24rB2)1/2\displaystyle\lambda+\frac{1}{L^{2}}\left(\left(1+r_{B}^{2}-r_{A}^{2}\right)^{2}-4r_{B}^{2}\right)^{-1/2} (87)

The second step 2\mathcal{I}_{2} is then another inversion across Y~c\tilde{Y}_{c}. Combining both steps, we obtain the required special conformal transformation 𝒦=21\mathcal{K}=\mathcal{I}_{2}\circ\mathcal{I}_{1}:

𝒦:Yμ(Yμaμ)bμ(Ya)212(Ya)b+b2(Ya)2,aμ=Ycμ,bμ=Y~cμYcμ\mathcal{K}:\;\;Y^{\mu}\to\frac{(Y^{\mu}-a^{\mu})-b^{\mu}(Y-a)^{2}}{1-2(Y-a)\cdot b+b^{2}(Y-a)^{2}},\;\;a^{\mu}=Y^{\mu}_{c},\;\;b^{\mu}=\tilde{Y}^{\mu}_{c}-Y^{\mu}_{c} (88)

and the radii of the co-centric images A~\tilde{A} and B~\tilde{B} of the two spheres become:

RA~=(1λ)2L2RA2RA,RB~=RB2λ2L2RBR_{\tilde{A}}=\frac{(1-\lambda)^{2}L^{2}-R_{A}^{2}}{R_{A}},\;\;R_{\tilde{B}}=\frac{R_{B}^{2}-\lambda^{2}L^{2}}{R_{B}} (89)

The symmetries of the conformal vacuum that preserves both spheres AA and BB can then be constructed using 𝒦\mathcal{K}:

^A,B(Ω)=𝒦1^(Ω)𝒦,R~(Ω)SO(d2)\hat{\mathcal{R}}_{A,B}(\Omega)=\mathcal{K}^{-1}\circ\hat{\mathcal{R}}(\Omega)\circ\mathcal{K},\;\;\tilde{R}(\Omega)\in\text{SO}(d-2) (90)

The natural modes of expansion ΦA,B(,m)\Phi^{(\ell,m)}_{A,B} on AA associated with the symmetry action ^A,B\hat{\mathcal{R}}_{A,B} can then be related to the usual spherical harmonics Φ(,m)\Phi^{(\ell,m)} associated with ordinary rotations ^\hat{\mathcal{R}} by pulling back the special conformal transformation:

ΦA,B(,m)=Φ(,m)𝒦\Phi^{(\ell,m)}_{A,B}=\Phi^{(\ell,m)}\circ\mathcal{K} (91)

where 𝒦:(α,ei)(α~,e~i)\mathcal{K}:(\alpha,e_{i})\to(\tilde{\alpha},\tilde{e}_{i}) now maps between angular variables on AA and A~\tilde{A}, whose form we can deduce by restricting (88) on the two spheres:

tanα~\displaystyle\tan{\tilde{\alpha}} =\displaystyle= RA(KRAcosα)K~RA2+K22RAKcosαK~2sin2α(K~sinα+RAcosαK)(K~RAcotα+Kcscα)\displaystyle\frac{R_{A}\left(K-R_{A}\cos{\alpha}\right)-\tilde{K}\sqrt{R_{A}^{2}+K^{2}-2R_{A}K\cos{\alpha}-\tilde{K}^{2}\sin^{2}{\alpha}}}{\left(\tilde{K}\sin{\alpha}+R_{A}\cos{\alpha}-K\right)\left(\tilde{K}-R_{A}\cot{\alpha}+K\csc{\alpha}\right)}
K~\displaystyle\tilde{K} =\displaystyle= (λλ~)[(1λ)2L2RA2]L\displaystyle(\lambda-\tilde{\lambda})\left[(1-\lambda)^{2}L^{2}-R_{A}^{2}\right]L
K\displaystyle K =\displaystyle= (1λ)L,e~i=ei\displaystyle(1-\lambda)L,\;\;\tilde{e}_{i}=e_{i} (92)

Furthermore, the definition of inner products between these modes ΦA,B(,m)\Phi^{(\ell,m)}_{A,B} come with an additional change of integration measures due to the conformal factor 𝒥(Ω)|𝒦|\mathcal{J}(\Omega)\sim|\partial\mathcal{K}|:

ΦA,B(,m),ΦA,B(,m)=Sd2A𝑑Ωd2𝒥(Ω)d2[ΦA,B(,m)(Ω)]×ΦA,B(,m)(Ω)\left\langle\Phi^{(\ell,m)}_{A,B},\;\Phi^{(\ell^{\prime},m^{\prime})}_{A,B}\right\rangle=\int_{S^{A}_{d-2}}d\Omega_{d-2}\mathcal{J}(\Omega)^{d-2}\left[\Phi^{(\ell,m)}_{A,B}(\Omega)\right]^{*}\times\Phi^{(\ell^{\prime},m^{\prime})}_{A,B}(\Omega) (93)

In our case, the deformation field is actually the contravariant radial component of a normal pointing vector field ζ(Ω)=ζr(Ω)\zeta(\Omega)=\zeta^{r}(\Omega), so its decomposition into harmonics modes using pull back (91) of the special conformal transformation 𝒦\mathcal{K} takes the form:

ζr(Ω)\displaystyle\zeta^{r}(\Omega) =\displaystyle= ,mζ(,m)[𝒦(Ω)1]r~r(Φ(,m)𝒦)(Ω)\displaystyle\sum_{\ell,m}\zeta^{(\ell,m)}\left[\partial\mathcal{K}(\Omega)^{-1}\right]^{r}_{\tilde{r}}\left(\Phi^{(\ell,m)}\circ\mathcal{K}\right)(\Omega) (94)
=\displaystyle= 𝒥1(α,ei),mζ(,m)ΦA,B(,m)(Ω)\displaystyle\mathcal{J}^{-1}(\alpha,e_{i})\sum_{\ell,m}\zeta^{(\ell,m)}\Phi^{(\ell,m)}_{A,B}(\Omega)

The full mapping (5) between angular variables simplifies in the OPE limit RB0R_{B}\to 0 that we have been working with so far:

tanα~(RA2L2)sinα(L2+RA2)cosα2LRA,e~i=ei\displaystyle\tan{\tilde{\alpha}}\to\frac{\left(R_{A}^{2}-L^{2}\right)\sin{\alpha}}{\left(L^{2}+R_{A}^{2}\right)\cos{\alpha}-2LR_{A}},\;\;\tilde{e}_{i}=e_{i} (95)

We can then obtain the conformal factor 𝒥\mathcal{J} as:

𝒥(α,ei)=1rA21+rA22rAcosα\mathcal{J}(\alpha,e_{i})=\frac{1-r_{A}^{2}}{1+r_{A}^{2}-2r_{A}\cos{\alpha}} (96)

which is what appears in our main result (58). At this stage, we can re-write (58) in terms of the above mode decomposition:

δIA,BΔ,=0𝑑Ωd2(1rA21+rA22rAcosα)d1ζ(Ω)\displaystyle\delta I^{\Delta,\ell=0}_{A,B}\propto\int d\Omega_{d-2}\left(\frac{1-r_{A}^{2}}{1+r_{A}^{2}-2r_{A}\cos{\alpha}}\right)^{d-1}\zeta(\Omega) (97)
=\displaystyle= ,mζ(,m)𝑑Ωd2𝒥(Ω)d2ΦA,B(,m)(Ω),mζ(,m)Φ(0,0),Φ(,m)=ζ(0,0)\displaystyle\sum_{\ell,m}\zeta^{(\ell,m)}\int d\Omega_{d-2}\mathcal{J}(\Omega)^{d-2}\;\Phi^{(\ell,m)}_{A,B}(\Omega)\propto\sum_{\ell,m}\zeta^{(\ell,m)}\left\langle\Phi^{(0,0)},\Phi^{(\ell,m)}\right\rangle=\zeta^{(0,0)}

We see that under an arbitrary shape deformation ζ(Ω)\zeta(\Omega), the linear response of mutual information in the OPE limit only picks up the zero-mode component ζ(0,0)\zeta^{(0,0)} of ζ(Ω)\zeta(\Omega) according to (94). In other words, the OPE limit of IA,BI_{A,B} extremizes against all non-zero modes of shape deformation:

δIA,BΔ,=0δζ(,m)=0,(,m)(0,0)\frac{\delta I^{\Delta,\ell=0}_{A,B}}{\delta\zeta^{(\ell,m)}}=0,\;\;\;(\ell,m)\neq(0,0) (98)

We can understand better the zero-mode nature of the response δIA,BΔ,=0ζ(0,0)\delta I^{\Delta,\ell=0}_{A,B}\propto\zeta^{(0,0)} from a symmetry point of view. When acting on the sphere AA, the zero-mode component ζ(0,0)\zeta^{(0,0)} of deformation:

ζ0,0(Ω)\displaystyle\zeta^{0,0}(\Omega) =\displaystyle= ζ(0,0)×𝒥1(α,ei)\displaystyle\zeta^{(0,0)}\times\mathcal{J}^{-1}\left(\alpha,e_{i}\right) (99)
RB0\displaystyle\underset{R_{B}\to 0}{\approx} ζ(0,0)(1+rA21rA2)ζ(0,0)(2rAcosα1rA2)\displaystyle\;\zeta^{(0,0)}\left(\frac{1+r_{A}^{2}}{1-r_{A}^{2}}\right)-\zeta^{(0,0)}\left(\frac{2r_{A}\cos{\alpha}}{1-r_{A}^{2}}\right)

implements simultaneously a dilation of RAR_{A} and a translation of the center YAY_{A} by:

RARA+ζ(0,0)(1+rA21rA2),YA=YAζ(0,0)(2rA1rA2)^,^=YBYA|YBYA|R_{A}\to R_{A}+\zeta^{(0,0)}\left(\frac{1+r_{A}^{2}}{1-r_{A}^{2}}\right),\;\;Y_{A}=Y_{A}-\zeta^{(0,0)}\left(\frac{2r_{A}}{1-r_{A}^{2}}\right)\hat{\mathcal{L}},\;\;\hat{\mathcal{L}}=\frac{Y_{B}-Y_{A}}{|Y_{B}-Y_{A}|} (100)

Using (88, 90) one can write down a one-parameter χ\chi family of invariant spheres under the group actions of ^A,B\hat{\mathcal{R}}_{A,B}, whose radius and center (χ,Yχ)(\chi,Y_{\chi}) we parametrize by:

Rχ=χ,Yχ=YALχ^R_{\chi}=\chi,\;\;Y_{\chi}=Y_{A}-L_{\chi}\hat{\mathcal{L}} (101)

They provide a ^A,B\hat{\mathcal{R}}_{A,B}-symmetric foliation of space. In the OPE limit RB0R_{B}\to 0, the spheres are defined by satisfying the relation (see Fig 9):

Refer to caption
Figure 9: The one-parameter χ\chi family of invariant spheres in the OPE limit RB0R_{B}\to 0
LχLχ2χ2=LL2RA2Lχ=L2(1rA2±(1rA2)2+4rχ2),rχ=χ/L\frac{L_{\chi}}{L_{\chi}^{2}-\chi^{2}}=\frac{L}{L^{2}-R_{A}^{2}}\;\rightarrow\;L_{\chi}=\frac{L}{2}\left(1-r_{A}^{2}\pm\sqrt{(1-r_{A}^{2})^{2}+4r_{\chi}^{2}}\right),\;\;r_{\chi}=\chi/L (102)

Zooming near the original sphere AA with:

χ=RA+(1+rA21rA2)δ,δRA\chi=R_{A}+\left(\frac{1+r_{A}^{2}}{1-r_{A}^{2}}\right)\delta,\;\;\delta\ll R_{A} (103)

and picking the (+) branch of (102) one can see that the mode of deformation mapping between these invariant spheres:

RARA+δ(1+rA21rA2),YAYAδ(2rA1rA2)^R_{A}\to R_{A}+\delta\left(\frac{1+r_{A}^{2}}{1-r_{A}^{2}}\right),\;\;Y_{A}\to Y_{A}-\delta\left(\frac{2r_{A}}{1-r_{A}^{2}}\right)\hat{\mathcal{L}} (104)

corresponds to precisely the zero-mode in (100). In other words, at leading order under shape deformation, the OPE limit of mutual information is only affected by modes that map between invariant spheres under ^A,B\hat{\mathcal{R}}_{A,B}.

6 Discussions

In this paper, we computed the linear response of mutual information IABI_{AB} between two spheres AA and BB under shape deformation on one of the spheres AA+ζA\to A+\zeta. The calculation is based on the OPE of the mutual information and the corresponding modular Hamiltonian in the limit RB0R_{B}\to 0. This enabled us to study the linear response of each OPE contribution of IABI_{AB} separately. The main result (58) of our calculation shows that for the possible dominant contribution from a primary scalar operator of conformal dimension Δ\Delta, its susceptibility under shape deformation is universal:

δIA,BΔ,=0h(Ω)δζ(Ω)|YΩYB|2(d1)\frac{\delta I^{\Delta,\ell=0}_{A,B}}{\sqrt{h(\Omega)}\delta\zeta(\Omega)}\propto|Y_{\Omega}-Y_{B}|^{-2(d-1)} (105)

in the sense that it is independent of the conformal dimension Δ\Delta. Regarding whether such universality could be spoiled in other possibilities of dominant OPE contributions, e.g. from primary operators with spins, we did a case study for the spin =1\ell=1 contribution and computed its linear response. It is found that the universal form of (58) persists in this case. It is therefore tempting to speculate that (58) might be valid for contributions from primary operators with arbitrary Δ\Delta and \ell, and thus describes the universal linear response of IABI_{AB} in the OPE limit. We provided a symmetry perspective to understand the implication of (58). For the CFT vacuum, there exists a group ^A,B\hat{\mathcal{R}}_{A,B} of symmetries isomorphic to the rotation group which leaves both the state and the sub-regions AA and BB invariant. The deformation vector fields ζ(Ω)\zeta(\Omega) therefore admit a natural mode decomposition associated with the symmetry group RA,BR_{A,B}:

ζ(Ω)=,mζ(,m)𝒥(Ω)1ΦA,B(,m)(Ω)\zeta(\Omega)=\sum_{\ell,m}\zeta^{(\ell,m)}\mathcal{J}(\Omega)^{-1}\Phi^{(\ell,m)}_{A,B}(\Omega) (106)

under which our result (58) takes the form: δζIABζ(0,0)\delta_{\zeta}I_{AB}\propto\zeta^{(0,0)}. This implies that for IABI_{AB} between spherical sub-regions AA and BB, its OPE limit extremizes over all deformations of AA that breaks the ^A,B\hat{\mathcal{R}}_{A,B} symmetry of the configuration. We end the paper by discussing a few points and propose some related future directions.

  • Order of integrations: additive structure in δIAB\delta I_{AB}

    Our calculation relied on the application of the entanglement first law and shape perturbation theory. Schematically, the linear response is extracted from contour integrals of the form:

    δS/δζ(x)zz0𝑑zTzz(z,z¯,x)H,(z0,x)A\delta S/\delta\zeta(x)\propto\oint_{z\sim z_{0}}dz\left\langle T_{zz}(z,\bar{z},x)H\right\rangle,\;\;(z_{0},x)\in\partial A (107)

    i.e. finding the residue of the simple pole (zz0)1(z-z_{0})^{-1} in the correlator between stress tensor and modular Hamiltonian. How the simple poles are actually produced could reveal a deeper aspect of the modular Hamiltonian. In the cases that one can deal with, the modular Hamiltonians are generally expressed as a finite dimensional integral of few-body local operators:

    H=𝑑s1,,dsnH(s1,,sn)𝑑sH(s)H=\int ds_{1},...,ds_{n}\;H\left(s_{1},...,s_{n}\right)\equiv\int d\vec{s}\;H\left(\vec{s}\right) (108)

    Such a form could be exact and is a result of symmetry, as in the case of spherical subregion in CFT vacuum; or in our case it could be the leading order contribution of the OPE expansion. In either scenario, the order of integrations is important for getting the correct answer. In particular, it means that schematically:

    δS/δζ𝑑z𝑑sTzz(z)H(s)𝑑s𝑑zTzz(z)H(s)\delta S/\delta\zeta\propto\oint dz\int d\vec{s}\left\langle T_{zz}(z)H\left(\vec{s}\right)\right\rangle\cancel{\propto}\int d\vec{s}\oint dz\left\langle T_{zz}(z)H\left(\vec{s}\right)\right\rangle (109)

    In other words, the integral (108) does not respond to shape deformations additively, i.e. the integrand does not directly produce simple poles in zz in (107), e.g. by OPE with TzzT_{zz}, whose residues are then integrated. One should take this as a manifestation about the non-local nature of the entanglement first law, even when the modular Hamiltonian appears as integrals of few-body local operators. The entanglement structure is organized non-locally, whose response to shape deformation is capture by (108) only collectivelly. Therefore, (109) should be the natural order of integration, and one can expect simple zz poles in (107) to emerge only after all the integrals of dsd\vec{s}.

    In this aspect, the OPE of ΔHAB\Delta H_{AB} responds to shape deformation in a slightly special way. The linear response is produced by the double integral term:

    ΔHAB𝑑sj𝑑skΔHAB(2)(sj,sk)\Delta H_{AB}\supset\int ds_{j}ds_{k}\;\Delta H^{(2)}_{AB}(s_{j},s_{k}) (110)

    in (3), where we recall:

    ΔHAB(2)(sj,sk)𝒪L(iskisj)𝒪L(isj)\Delta H^{(2)}_{AB}(s_{j},s_{k})\propto\mathcal{O}_{L}(-is_{k}-is_{j})\mathcal{O}_{L}(-is_{j}) (111)

    In this case, one of the integrals (in sjs_{j}) is sufficient to produce simple poles in zz. This then allows us to partially switch the order of integration and write:

    δIAB𝑑z(𝑑sj𝑑skTzz(z)ΔHAB(sj,sk))\displaystyle\delta I_{AB}\propto\oint dz\left(\int ds_{j}ds_{k}\left\langle T_{zz}(z)\Delta H_{AB}(s_{j},s_{k})\right\rangle\right) (112)
    =\displaystyle= 𝑑sk(𝑑z𝑑sjTzz(z)ΔHAB(sj,sk))=𝑑sk``δIAB(sk)"\displaystyle\int ds_{k}\left(\oint dz\int ds_{j}\left\langle T_{zz}(z)\Delta H_{AB}(s_{j},s_{k})\right\rangle\right)=\int ds_{k}\;``\delta I_{AB}(s_{k})"

    We interpret this as follows. In terms of response to shape deformation, the modular Hamiltonian (3) contributes additively in sks_{k} which measures the relative modular flow between the bi-local operator 𝒪(isjisk)𝒪(isj)\mathcal{O}(-is_{j}-is_{k})\mathcal{O}(-is_{j}) in the integrand. One is then naturally led to wonder whether the additive behavior of sks_{k} bears more insights into the entanglement structures relevant for mutual information in the OPE limit. In the future, it would be interesting to examine this deeper, for example by probing the additive behavior from other perspectives.

  • Universality of (58) among all OPE contributions?

    The main result (58) of the paper exhibits a high degree of universality. Its form factor only depends on the space-time dimension dd. Furthermore, one obtains the same form factor when the primary operator 𝒪\mathcal{O} forming the bi-local OPE contribution is a scalar or have spin =1\ell=1. We had not planned to perform an extensive check of the validity of (58) among all OPE contributions. This in particular could include:

    1. 1.

      bi-local operators formed by higher spin 2\ell\geq 2 primary operators;

    2. 2.

      bi-local contributions from descendent operators;

    3. 3.

      multi-local channel OPE contributions of the form:

      ΣBnci1,,in𝒪(i1)𝒪(in)\Sigma^{n}_{B}\rightarrow c_{i_{1},...,i_{n}}\mathcal{O}^{(i_{1})}...\mathcal{O}^{(i_{n})}

    In retrospect, the form factor in (58) could have been fixed by symmetry if we accept the following two points regarding the linear responses: (i) it transforms covariantly under conformal transformations; (ii) in the conformal frame where A~\tilde{A} and B~\tilde{B} are co-centric (see section 5), it is rotationally invariant on A~\tilde{A}. The second point is the dynamical input that has to be provided by our calculations, but otherwise it is a natural reflection of the underlying symmetry, which may also explain the universality featured in (58). On the other hand, at the level of actual computations we have not found simple organizing principles making the universality transparent – the agreement between =1\ell=1 and scalar contributions currently appears as a result of many cancellations among terms violating (58). For the case of multi-local channels, the structure of the computations becomes more obscure: (i) the contribution to the mutual information itself would be more theory-dependent, e.g. depending on the OPE coefficient COOOC_{OOO}; (ii) the shape dependence involves higher-point correlation functions Tμν𝒪𝒪\sim\langle T_{\mu\nu}\mathcal{O}...\mathcal{O}\rangle that cannot be fixed by symmetries; (iii) it is not clear how much of the shape-dependence can be constrained by consistency with (60), i.e. conformal ward identity. Putting these difficulties and uncertainties aside, if the result is truly universal, this may point to more hidden structures in the perturbative expansion of the entanglement first law governing the form of linear responses. Therefore, in the future it is interesting to further study the validity of, or otherwise modifications to (58) from more generic contributions, e.g. those listed in (1-3).

  • Beyond the OPE limit: finite RBR_{B} corrections

    In this paper, we focused on the OPE limit of mutual information. In the conformal frame of our interest, this corresponds to the point-like limit: RBL,RAR_{B}\ll L,R_{A}. The result (58) pertains only to the leading order contribution in this limit. In the future, it is important to extend the result beyond the OPE limit. This could be achieved by re-summing over the responses from all OPE contributions. Similar to ordinary OPE expansion, these contributions can be organized into conformal blocks associated with particular primary operators in the orbifold CFT (e.g. bi-local or multi-local channels in 1-3). They include corrections from descendent operators, which are essentially derivative expansions that restore the full shape dependence on sub-region BB. As a result, studying the shape dependence of mutual information at the level of conformal blocks will be an important future direction to pursue that can systematically push beyond the OPE limit future_work . With the possibility of universal multiplicative form factor in front of each OPE contribution, it is plausible that after re-summation over descendent corrections, such feature survives and the linear response from each conformal block contribution is still encoded in a multiplicative form factor. In fact, the symmetry of the set-up is independent of the OPE limit – the subregion-preserving symmetry group ^AB\hat{\mathcal{R}}_{AB} exists for arbitrary RA,RBR_{A},R_{B}. Therefore, if the unversality and extremality are also valid beyond the OPE limit, the susceptibility for finite RBR_{B} is simply given by:

    δIABh(Ω)δζ(Ω)G(α)(1+G(α)2)1d2[G(α)sin(α)]d3\frac{\delta I_{AB}}{\sqrt{h(\Omega)}\delta\zeta(\Omega)}\propto G^{\prime}(\alpha)\Big{(}1+G(\alpha)^{2}\Big{)}^{\frac{1-d}{2}}\left[\frac{G(\alpha)}{\sin{(\alpha)}}\right]^{d-3} (113)

    where G(α)G(\alpha) is the complicated function of α\alpha that appears on the right-hand side of the first line in (5). In the future, it is interesting to check if this is indeed the case.

  • Higher order responses

    Last but not least, it is also important to extend our analysis of shape dependence beyond the linear order response in shape deformation. A natural next-step towards this is to compute the non-local quadratic order response future_work :

    δ2IABδζ(x1)ζ(x2),x1x2\frac{\delta^{2}I_{AB}}{\delta\zeta(x_{1})\zeta(x_{2})},\;\;x_{1}\neq x_{2} (114)

    An analogous quantity for the entanglement entropy on spheres (called entanglement density) were studied in Faulkner2016Apr , and were found to encode universal data of the underlying CFTs. Similarly, we expect to learn more about CFT and its entanglement structures from a corresponding second order response of mutual information. For example, from a shape-dependence point of view, the strong sub-additivity constraint for mutual information Casini2021Sep bares no content at linear order – it is trivially saturated by the linear order responses due to its Markovian form. The constraint becomes meaningful only at the second order. In the face of extremality property of mutual information between spheres, we also need to push to the second order response to identify the nature of the extremum, i.e. is it a maximum or a minimum? It will be particularly interesting to study higher order responses in conjunction with higher order corrections in small RBR_{B}. We leave these for future investigations.

Acknowledgments

We thank Thomas Faulkner, Ling-Yan Hung, Shenghan Jiang, and Xinan Zhou for useful discussions. We thank Mark Mezei and Gonzalo Torroba for comments on the manuscript. This work is supported by National Science Foundation of China (NSFC) grant no. 12175238.

Appendix A Computing the integrals

In this appendix, we will give some of the details for computing the integrals that we encounter when we try to calculate the shape dependence of mutual information. They take the following form:

δIA,Bδξ(Xi)\displaystyle\frac{\delta I_{A,B}}{\delta\xi(X^{i})} =(i)|z|=ϵ~𝑑z(Tzz(z,z¯,Xi)+Tzz¯(z,z¯,Xi))ΔH~A,B+h.c\displaystyle=(i)\oint_{|z|=\tilde{\epsilon}}{d}z\,\ \left\langle\bigg{(}T_{zz}(z,\bar{z},X^{i})+T_{z\bar{z}}(z,\bar{z},X^{i})\bigg{)}\Delta\tilde{H}_{A,B}\right\rangle+h.c (115)
=𝒦(1)+𝒦(2)\displaystyle=\mathcal{K}^{(1)}+\mathcal{K}^{(2)}

where 𝒦(1),𝒦(2)\mathcal{K}^{(1)},\mathcal{K}^{(2)} represent contributions from the single and double integral terms of ΔH~A,B\Delta\tilde{H}_{A,B}. They take the explicit form:

𝒦(1)\displaystyle\mathcal{K}^{(1)} =iρ2Δ|z|=ϵ~dzds(2cosh(s/2))2Δ+2(Tzz(z,z¯,Xi)+Tzz¯(z,z¯,Xi))\displaystyle=-i\rho^{2\Delta}\oint_{|z|=\tilde{\epsilon}}{d}z\,\ \int_{-\infty}^{\infty}\frac{ds}{(2\cosh(s/2))^{2\Delta+2}}\bigg{\langle}\bigg{(}T_{zz}(z,\bar{z},X^{i})+T_{z\bar{z}}(z,\bar{z},X^{i})\bigg{)} (116)
×𝒪(X[s])𝒪(X[0])+h.c\displaystyle\times\mathcal{O}\left(-X[s]\right)\mathcal{O}\left(X[0]\right)\bigg{\rangle}+h.c
𝒦(2)\displaystyle\mathcal{K}^{(2)} =ρ2Δ2π|z|=ϵ~𝑑zdsjdsk4cosh2(sj/2)(1esk+iϵ1+1esk+sj+1)c1(isk+ϵ)\displaystyle=-\frac{\rho^{2\Delta}}{2\pi}\oint_{|z|=\tilde{\epsilon}}{d}z\,\ \int_{-\infty}^{\infty}\frac{ds_{j}ds_{k}}{4\cosh^{2}(s_{j}/2)}\left(\frac{1}{e^{s_{k}+i\epsilon}-1}+\frac{1}{e^{s_{k}+s_{j}}+1}\right)c_{1}(-is_{k}+\epsilon)
×(Tzz(z,z¯,Xi)+Tzz¯(z,z¯,Xi))𝒪(X[sj+sk])𝒪(X[sj])+h.c.\displaystyle\times\left\langle\bigg{(}T_{zz}(z,\bar{z},X^{i})+T_{z\bar{z}}(z,\bar{z},X^{i})\bigg{)}\mathcal{O}\left(-X[s_{j}+s_{k}]\right)\mathcal{O}\left(-X[s_{j}]\right)\right\rangle+h.c.

where we recall that X[s]±=e±s,X[s]i=0X[s]^{\pm}=e^{\pm s},X[s]^{i}=0. For simplicity, we only focus on the 𝑑z\oint{dz} integrals. The main task is to extract the residue of simple poles in zz from the integrands. Furthermore, using the analysis to be presented below it can be checked that only the correlators involving TzzT_{zz} can give rise to such poles of zz, so we shall omit writing terms involving other components of stress tensor from now to declutter notations. The three-point functions TzzOO\left\langle T_{zz}OO\right\rangle by themselves do not exhibit simple poles in zz. Similar to (20) such poles can only emerge as a result of performing the modular integrals, i.e. we should identify the following behaviors from the integrands:

(1)(z,z¯)\displaystyle\mathcal{F}^{(1)}(z,\bar{z}) =\displaystyle= 𝑑sk1(s)Tzz(z,z¯,Xi)𝒪(X[s])𝒪(X[0])z+\displaystyle\int ds\;k_{1}(s)\left\langle T_{zz}\left(z,\bar{z},X^{i}\right)\mathcal{O}\left(-X[s]\right)\mathcal{O}\left(X[0]\right)\right\rangle\sim\frac{\sharp}{z}+...
(2)(z,z¯)\displaystyle\mathcal{F}^{(2)}(z,\bar{z}) =\displaystyle= 𝑑sj𝑑skk2(sj,sk)Tzz(z,z¯,Xi)𝒪(X[sj+sk])𝒪(X[sj])z+\displaystyle\int ds_{j}ds_{k}\;k_{2}\left(s_{j},s_{k}\right)\left\langle T_{zz}\left(z,\bar{z},X^{i}\right)\mathcal{O}\left(-X[s_{j}+s_{k}]\right)\mathcal{O}\left(-X[s_{j}]\right)\right\rangle\sim\frac{\sharp}{z}+...

for the terms of 𝒦(1)\mathcal{K}^{(1)} and 𝒦(2)\mathcal{K}^{(2)} respectively. Doing these integrals exactly is neither pleasant nor illuminating. However, since we are only interested in the singularity of (1,2)(z,z¯)\mathcal{F}^{(1,2)}(z,\bar{z}) as z,z¯0z,\bar{z}\to 0, it is revealing to study their nature of divergences when setting z,z¯z,\bar{z} to be exactly zero. For example, a necessary condition for the emergence of simple pole is that (1,2)(0,0)\mathcal{F}^{(1,2)}(0,0) be divergent. This will rule out the possibility of finding poles in (1)(z,z¯)\mathcal{F}^{(1)}(z,\bar{z}) which implies that 𝒦(1)=0\mathcal{K}^{(1)}=0. To check this, let us quote the three-point functions (52) and compute (1)(0,0)<\mathcal{F}^{(1)}(0,0)<\infty explicitly as:

(1)(0,0)\displaystyle\mathcal{F}^{(1)}(0,0) =\displaystyle= ds(2cosh(s/2))2Δ+2Tzz(0,0;Xi)𝒪(X[s])𝒪(X[0])\displaystyle\int^{\infty}_{-\infty}\frac{ds}{\left(2\cosh{(s/2)}\right)^{2\Delta+2}}\left\langle T_{zz}(0,0;X^{i})\mathcal{O}(-X[s])\mathcal{O}(X[0])\right\rangle (117)
=\displaystyle= C12T4(1+D)(1+d)0duu2Δd/2(u+1)4Δ+3d,ues\displaystyle-\frac{C_{12T}}{4}\left(1+D\right)^{-(1+d)}\int^{\infty}_{0}\frac{du\;u^{2\Delta-d/2}}{(u+1)^{4\Delta+3-d}},\;\;u\equiv e^{s}
=\displaystyle= (C12Tπ24Δd+4)(1+D)(1+d)Γ(2Δd2+1)Γ(2Δd2+32)<\displaystyle-\left(\frac{C_{12T}\sqrt{\pi}}{2^{4\Delta-d+4}}\right)\left(1+D\right)^{-(1+d)}\frac{\Gamma\left(2\Delta-\frac{d}{2}+1\right)}{\Gamma\left(2\Delta-\frac{d}{2}+\frac{3}{2}\right)}<\infty

where we have denoted D=i=2d1(Xi)2D=\sum_{i=2}^{d-1}(X^{i})^{2}.

So we are only left with the possible contribution from 𝒦(2)\mathcal{K}^{(2)}, whose integrand is given by quoting the corresponding three-point function as:

(2)(z,z¯)=𝑑sj𝑑skk2(sj,sk)Tzz(z,z¯;Xi)𝒪(X[sj+sk])𝒪(X[sj])\displaystyle\mathcal{F}^{(2)}(z,\bar{z})=\int^{\infty}_{-\infty}ds_{j}\int^{\infty}_{-\infty}ds_{k}\;k_{2}\left(s_{j},s_{k}\right)\left\langle T_{zz}(z,\bar{z};X^{i})\mathcal{O}\left(-X[s_{j}+s_{k}]\right)\mathcal{O}\left(-X[s_{j}]\right)\right\rangle (118)
=\displaystyle= C12T4𝑑sk𝑑sjk2(sj,sk)|w12|d22Δ(w12(z¯w¯1)(z¯w¯2)w¯12D)2((zw1)(z¯w¯1)+D)1+d2((zw2)(z¯w¯2)+D)1+d2\displaystyle\frac{C_{12T}}{4}\int^{\infty}_{-\infty}ds_{k}\int^{\infty}_{-\infty}ds_{j}\frac{k_{2}(s_{j},s_{k})\;\left|w_{12}\right|^{d-2-2\Delta}\Big{(}w_{12}(\bar{z}-\bar{w}_{1})(\bar{z}-\bar{w}_{2})-\bar{w}_{12}D\Big{)}^{2}}{\Big{(}(z-w_{1})(\bar{z}-\bar{w}_{1})+D\Big{)}^{1+\frac{d}{2}}\Big{(}(z-w_{2})(\bar{z}-\bar{w}_{2})+D\Big{)}^{1+\frac{d}{2}}}
=\displaystyle= (1)d2(C12T4)0𝑑v(v1)d2v1d20du(u+1)(uv+1)\displaystyle(-1)^{\frac{d}{2}}\left(\frac{C_{12T}}{4}\right)\int^{\infty}_{0}dv\;(v-1)^{d-2}v^{1-\frac{d}{2}}\int^{\infty}_{0}\frac{du}{(u+1)(uv+1)}
×\displaystyle\times (z¯2u+(v1+1)z¯+u1v1(1+D))2(uvz¯+u1v1z+1+D)1+d2(uz¯+u1z+1+D)1+d2\displaystyle\frac{\left(\bar{z}^{2}u+(v^{-1}+1)\bar{z}+u^{-1}v^{-1}(1+D)\right)^{2}}{\left(uv\bar{z}+u^{-1}v^{-1}z+1+D\right)^{1+\frac{d}{2}}\left(u\bar{z}+u^{-1}z+1+D\right)^{1+\frac{d}{2}}}

where the coordinates (w1,2,w¯1,2)(w_{1,2},\bar{w}_{1,2}) are given by w1=esk+sj,w¯1=esksj,w2=esj,w¯2=esjw_{1}=-e^{s_{k}+s_{j}},\bar{w}_{1}=-e^{-s_{k}-s_{j}},w_{2}=-e^{s_{j}},\bar{w}_{2}=-e^{-s_{j}} and w12=w1w2w_{12}=w_{1}-w_{2}, etc. In the last step we have re-defined the modular parameters in terms of u=esj,v=esku=e^{s_{j}},v=e^{s_{k}} and discarded the |z|2=ϵ~21|z|^{2}=\tilde{\epsilon}^{2}\ll 1 terms in the denominator. In this case, one can easily check that (2)(0,0)\mathcal{F}^{(2)}(0,0) is divergent from the uu integral alone:

(2)(0,0)0duu2v2(u+1)(uv+1)(1+D)dlimu0(u1)=\mathcal{F}^{(2)}(0,0)\propto\int^{\infty}_{0}\frac{du\;u^{-2}v^{-2}}{(u+1)(uv+1)}(1+D)^{-d}\sim\lim_{u\to 0}\left(u^{-1}\right)=\infty (119)

We can extract from the nature of divergence (119) two pieces of information relevant for our purposes. Firstly the divergence is controlled by the u0u\to 0 limit of the integral; secondly the order of the divergence dictates that (2)(z,z¯)\mathcal{F}^{(2)}(z,\bar{z}) has a simple pole for zz:

limz0(2)(z,z¯)z+\lim_{z\to 0}\mathcal{F}^{(2)}(z,\bar{z})\to\frac{\sharp}{z}+... (120)

This is because any finite value of zz provides an effective regulator for the u0u\to 0 divergence through the u1v1zu^{-1}v^{-1}z and u1zu^{-1}z terms in the denominator of (118). Therefore we can focus on the u0u\to 0 regime of the uu integral and thus neglect the z¯\bar{z} terms in the denominator of (118). Under this approximation, the relevant part of the integrand become a holomorphic function in zz 111The treatment for the 𝑑z¯\oint d\bar{z} integral of the correlators involving Tz¯z¯T_{\bar{z}\bar{z}} proceeds in exactly analogous way in terms of anti-holomorphic functions.. It is interesting to see that we do not need to complete both integral in uu and vv in order to reveal the simple poles in zz, only the uu integral alone is sufficient. The final result is then obtained by integrating the vv-dependent residue over the remaining modular parameter vv. By doing this we have partially switched the order of integration between 𝑑z\oint dz and 𝑑v\int dv. As is discussed in section 6, the fact that this works implies that the shape response exhibits additive behavior in the vv variable of the modular Hamiltonian, and may point to deeper aspects regarding the underlying structures of modular Hamiltonian in OPE.

Now we lay out more details for the computations. We first focus on the uu-integral of (2)\mathcal{F}^{(2)} and extract the residue of poles in zz. We will be slightly more general than the specific form of (118), since a similar procedure can be applied to computing contributions from spinning primary operator like 𝒪μ\mathcal{O}_{\mu} in section (4.4), whose numerators from the three-point functions Tzz𝒪μ𝒪ν\left\langle T_{zz}\mathcal{O}_{\mu}\mathcal{O}_{\nu}\right\rangle are more complicated functions of u,v,z,z¯u,v,z,\bar{z}. Stripping off the uu-independent factors, a generic uu-integral term takes the following form:

0𝑑u(u+1)1(uv+1)1ud+n(uv+z/(1+D))1+d/2(u+z/(1+D))1+d/2,n\displaystyle\int_{0}^{\infty}{d}u\,\ \frac{(u+1)^{-1}(uv+1)^{-1}u^{d+n}}{\Big{(}uv+z/(1+D)\Big{)}^{1+d/2}\Big{(}u+z/(1+D)\Big{)}^{1+d/2}},\;\;n\in\mathds{Z} (121)
=\displaystyle= Γ(2+d)Γ(1+d/2)201𝑑wwd/2(1w)d/20du(u+1)1(uv+1)1ud+n(uvw+u(1w)+z/(1+D))2+d\displaystyle\frac{\Gamma(2+d)}{\Gamma(1+d/2)^{2}}\int_{0}^{1}{d}w\;w^{d/2}(1-w)^{d/2}\int_{0}^{\infty}\frac{du\;(u+1)^{-1}(uv+1)^{-1}u^{d+n}}{\Big{(}uvw+u(1-w)+z/(1+D)\Big{)}^{2+d}}\;\;

where we have used the Feymann parameter technique to further simplify the uu integral part. Now we can extract the singularity in z0z\to 0 of (121) by defining:

y=u+δ,δ=z(1+D)1(wv+1w)11y=u+\delta,\;\;\delta=z(1+D)^{-1}(wv+1-w)^{-1}\ll 1 (122)

and zooming into the uδu\sim\delta region of the uu integral in (121):

0𝑑u(u+1)1(uv+1)1ud+n(u(wv+1w)+z/(1+D))d+2\displaystyle\int^{\infty}_{0}d{u}\,\ \frac{(u+1)^{-1}(uv+1)^{-1}u^{d+n}}{\Big{(}u(wv+1-w)+z/(1+D)\Big{)}^{d+2}} (123)
=\displaystyle= 1(wv+1w)d+2(δ𝑑yy(d+2)(yδ)d+n)(1+𝒪(δ))\displaystyle\frac{1}{\left(wv+1-w\right)^{d+2}}\left(\int^{\infty}_{\delta}dy\,\ y^{-(d+2)}\left(y-\delta\right)^{d+n}\right)\left(1+\mathcal{O}\left(\delta\right)\right)
=\displaystyle= 1(wv+1w)d+1+nΓ(1n)Γ(d+1+n)Γ(d+2)(z1+D)1+n(1+𝒪(z1+D))\displaystyle\frac{1}{\left(wv+1-w\right)^{d+1+n}}\frac{\Gamma(1-n)\Gamma(d+1+n)}{\Gamma(d+2)}\left(\frac{z}{1+D}\right)^{-1+n}\left(1+\mathcal{O}\left(\frac{z}{1+D}\right)\right)

For the scalar contribution (118) we only have one term with n=0n=0, which corresponds to a simple pole in zz. For the spin-1 contribution in (4.4) there are terms with un,n<0u^{n},n<0, but they are multiplied additional powers of zz to still give at most simple poles in zz 222If there are higher order poles in zz, our result will receive contributions that depend on non-universal details of the cut-off tube, e.g. shape of the tube which are higher-order corrections in small ϵ\epsilon.. Setting n=0n=0 from now on, one can then collect the residue of the simple zz-pole in (123), feed into the remaining ww integral in (121) and obtain a hypergeometric function in 1v1-v. We shall not write down these intermediate details. Finally, one ends up with the following vv-integral for 𝒦(2)\mathcal{K}^{(2)}:

𝒦(2)\displaystyle\mathcal{K}^{(2)} =(2iρ2Δd+1)(C12T4)(1+D)d+1(1)d/212Δ\displaystyle=-\left(\frac{2i\rho^{2\Delta}}{d+1}\right)\left(\frac{C_{12T}}{4}\right)(1+D)^{-d+1}(-1)^{d/2-1-2\Delta} (124)
×0dvv2Δ(v1)4Δ+1dF12(1+d,1+d2;2+d;1v)\displaystyle\times\int_{0}^{\infty}{d}v\,\ \frac{v^{2\Delta}}{(v-1)^{4\Delta+1-d}}\,\ {{}_{2}F_{1}}\bigg{(}1+d,1+\frac{d}{2};2+d;1-v\bigg{)}

The hypergeometric function in the integrand admits a finite expansion for integer values of dd. For example, in the case of d=4d=4, it takes the form:

F12(1+d,1+d2;2+d;1v)=20v1(v1)55v22(v1)520v(v1)5+5v22(v1)5+30logv(v1)5\displaystyle{{}_{2}F_{1}}\Big{(}1+d,1+\frac{d}{2};2+d;1-v\Big{)}=\frac{20v^{-1}}{(v-1)^{5}}-\frac{5v^{-2}}{2(v-1)^{5}}-\frac{20v}{(v-1)^{5}}+\frac{5v^{2}}{2(v-1)^{5}}+\frac{30\log{v}}{(v-1)^{5}}

The first four terms cancel in the integral due to the invariance of integrand under vv1v\to v^{-1}, so we only need to keep the logv\log{v} term. This pattern turns out to be valid for general even integer dd, whose logv\log{v} terms are given by:

F12(1+d,1+d2;2+d;1v)\displaystyle{{}_{2}F_{1}}\Big{(}1+d,1+\frac{d}{2};2+d;1-v\Big{)} =(1)d/2Γ(2+d)Γ(1+d/2)2log(v)(v1)d+1+\displaystyle=(-1)^{d/2}\frac{\Gamma(2+d)}{\Gamma(1+d/2)^{2}}\frac{\log(v)}{(v-1)^{d+1}}+\cdots (125)

Plugging this into the vv integral, we eventually obtain that:

𝒦(2)\displaystyle\mathcal{K}^{(2)} =2Δρ2Δπ42Δ+1Γ(2Δ+1)Γ(2Δ+3/2)2d2Γ(d12)πd/21/2(1+D)d+1.\displaystyle=2\Delta\rho^{2\Delta}\frac{\sqrt{\pi}}{4^{2\Delta+1}}\frac{\Gamma(2\Delta+1)}{\Gamma(2\Delta+3/2)}\frac{2^{d-2}\Gamma\left(\frac{d-1}{2}\right)}{\pi^{d/2-1/2}}(1+D)^{-d+1}. (126)

For odd integer dd the analysis is slightly more involved, but in the end one obtains the same dependence on dd, and we shall spare the details.

References