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

aainstitutetext: Department of Physics, National Tsing Hua University, Hsinchu 30013, Taiwanbbinstitutetext: Yukawa Institute for Theoretical Physics, Kyoto University, Kyoto 606-8502, Japanccinstitutetext: Department of Physics and Photon Science, Gwangju Institute of Science and Technology,
Gwangju 61005, Korea

A note on the non-planar corrections for the Page curve in the PSSY model via the IOP matrix model correspondence

Norihiro Iizuka c    and Mitsuhiro Nishida [email protected] [email protected]
Abstract

We develop a correspondence between the PSSY model and the IOP matrix model by comparing their Schwinger-Dyson equations, Feynman diagrams, and parameters. Applying this correspondence, we resum specific non-planar diagrams involving crossing in the PSSY model by using a non-planar analysis of a two-point function in the IOP matrix model. We also compare them with Page’s formula on entanglement entropy and discuss the contributions of extra-handle-in-bulk diagrams.

1 Introduction

Understanding the evaporation process of black holes Hawking:1975vcx has played an important role in our understanding of quantum gravity. Since quantum gravity is a challenging research subject, the approach of studying toy models of quantum gravity is beneficial. It is essential to investigate the quantum effects in such toy models where we can compute quantum correction exactly. The Penington-Shenker-Stanford-Yang (PSSY) model Penington:2019kki , sometimes called the West Coast model, is a very nice toy model of evaporating black holes. It is a 2-dimensional Jackiw-Teitelboim (JT) gravity, with end-of-the-world (EOW) branes. This model has two subsystems, a black hole with its Hilbert space dimension eSe^{\textbf{S}} represented by the flavors of the EOW branes, and an auxiliary reference system \mathbb{R} representing “radiation” with its Hilbert space dimension kk. It was shown in Penington:2019kki that depending on k<eSk<e^{\textbf{S}} which corresponds to early black hole or k>eSk>e^{\textbf{S}} which corresponds to late black hole, the dominant topology in the gravitational path integral changes and that leads to the Page curve behavior changes before and after the Page point Page:1993df ; Page:1993wv .

The analysis done by Penington:2019kki above is in the planar limit and to see the essential Page curve behavior change at the Page point, this planar analysis is good enough. However, it is certainly interesting to investigate the non-planar corrections to this model, which corresponds to the quantum gravity effects. The motivation of this paper is to investigate these non-planar corrections to the PSSY model.

One of the main results of this paper is to show that there is a curious correspondence between the PSSY model and the IOP matrix model Iizuka:2008eb , another toy model investigated before as a toy model of proving a black hole. The IOP matrix model is a cousin of the IP model Iizuka:2008hg and represents the decay of the correlation function of the probe fundamental field interacting with a matrix degree of freedom describing a black hole. The pros and cons of the IOP matrix model are that it is simpler than the IP model and thus one can solve it in various ways, but the correlator decays only by the power law, not by the exponential. As we will show explicitly, the correspondence is seen through the Feynman diagrams of both models. In the IOP matrix model, not only the planar contributions but also the leading non-planar contributions are explicitly calculated in Iizuka:2008eb . We show that using the correspondence between the PSSY model and the IOP matrix model, one can evaluate the specific non-planar corrections exactly in the PSSY model. This is the main point of this paper.

However, this is not the end of the story. In the PSSY model, in fact, there are two types of non-planar corrections. The correspondence between the PSSY model and the IOP matrix model enables us to evaluate only one class of non-planar corrections, which involves the diagrams of “crossing”. On the other hand, the other non-planar corrections are associated with the extra-handle-in-bulk diagrams. For the extra-handle-in-bulk diagrams, there is no associated diagram in the IOP matrix model. Thus, the correspondence is not completely one-to-one and it does not directly answer all non-planar corrections. Therefore, one needs to do the direct bulk calculation of resummation of such extra-handle-in-bulk diagrams. We leave these extra-handle-in-bulk calculations for future work.

The organization of this short note is as follows. In section 2, we review both the PSSY model and the IOP matrix model. Section 3 is our main result, where we show there is a correspondence between the PSSY model and the IOP matrix mode, and through that, we evaluate the non-planar corrections involving the diagram of crossing in the PSSY model. In section 4, we conclude and discuss open issues as well as possible generalizations of our works.

2 The PSSY model and the IOP matrix model in the planar limit

In this section, the PSSY model (or the West Coast model) and the IOP matrix model are reviewed. We focus on the spectral density of a reduced density matrix in a microcanonical ensemble of the PSSY model and the spectral density of a two-point function of fundamental fields in the IOP matrix model. After reviewing the two models, in Section 3, we will point out that both spectral densities in the planar limit are represented by the Marchenko-Pastur distribution in random matrix theory and explain how both models correspond to each other.

2.1 Review of the PSSY model

The PSSY model Penington:2019kki consists of a black hole in JT gravity with an end-of-the-world (EOW) brane behind the horizon with tension μ0\mu\geq 0. Its Euclidean action is

S=\displaystyle S= SJT+μBrane𝑑s,\displaystyle\;S_{\text{JT}}+\mu\int_{\text{Brane}}ds, (1)
SJT=\displaystyle S_{\text{JT}}= S04π(gR+2MhK)(12gϕ(R+2)+hϕK).\displaystyle\;-\frac{S_{0}}{4\pi}\left(\int_{\mathcal{M}}\sqrt{g}R+2\int_{\partial M}\sqrt{h}K\right)-\left(\frac{1}{2}\int_{\mathcal{M}}\sqrt{g}\phi(R+2)+\int_{\partial\mathcal{M}}\sqrt{h}\phi K\right). (2)

We impose the standard asymptotic boundary condition

ds2|=dτ2zϵ2,ϕ|=1zϵ,\displaystyle ds^{2}|_{\partial\mathcal{M}}=\frac{d\tau^{2}}{z_{\epsilon}^{2}},\;\;\;\phi|_{\partial\mathcal{M}}=\frac{1}{z_{\epsilon}}, (3)

where τ\tau is the boundary Euclidean time, and zϵz_{\epsilon} is the near-boundary cutoff.

Suppose that there are kk orthogonal states |i|i\rangle_{\mathbb{R}} of the “radiation” system \mathbb{R}, which are entangled with kk interior of the EOW brane microstates |ψi𝔹|\psi_{i}\rangle_{\mathbb{B}} of the black hole 𝔹\mathbb{B}. A pure state |Ψ|\Psi\rangle representing this entanglement is given by

|Ψ=1ki=1k|ψi𝔹|i,\displaystyle|\Psi\rangle=\frac{1}{\sqrt{k}}\sum\limits_{i=1}^{k}\,\ket{\psi_{i}}_{\mathbb{B}}\ket{i}_{\mathbb{R}}, (4)

where the radiation system \mathbb{R} can be interpreted as the early radiation of an evaporating black hole. The reduced density matrix ρ\rho_{\mathbb{R}} and its resolvent R(λ)R(\lambda) are defined by

ρ\displaystyle\rho_{\mathbb{R}} :=Tr𝔹|ΨΨ|=1ki,j=1k|ji|ψi|ψj𝔹,\displaystyle:=\Tr_{\mathbb{B}}\ket{\Psi}\bra{\Psi}=\frac{1}{k}\,\sum\limits_{i,j=1}^{k}\,\ket{j}\bra{i}_{\mathbb{R}}\,\braket{\psi_{i}}{\psi_{j}}_{\mathbb{B}}, (5)
R(λ)\displaystyle R(\lambda) :=i=1kRii(λ),Rij(λ):=(1λ𝟙ρ)ij=1λδij+n=11λn+1(ρn)ij.\displaystyle:=\sum\limits_{i=1}^{k}R_{ii}(\lambda)\,,\;\;\;R_{ij}(\lambda):=\left(\frac{1}{\lambda\mathds{1}-\rho_{\mathbb{R}}}\right)_{ij}\,=\,\frac{1}{\lambda}\,\delta_{ij}+\sum\limits_{n=1}^{\infty}\,\frac{1}{\lambda^{n+1}}\,(\rho_{\mathbb{R}}^{n})_{ij}\,. (6)
Refer to caption
Figure 1: Schwinger-Dyson equation for the PSSY model in the planar limit.

When e𝕊e^{\mathbb{S}}, which is the dimensions of 𝐁\bf{B}, and kk, which is the dimension of 𝐑\bf{R}, are large, only planar diagrams are dominant in the Schwinger-Dyson equation of Rij(λ)R_{ij}(\lambda), as Fig. 1, thus we obtain

Rij(λ)=1λδij+1λn=1ZnDisk(kZ1Disk)nR(λ)n1Rij(λ),\displaystyle R_{ij}(\lambda)=\,\frac{1}{\lambda}\,\delta_{ij}+\frac{1}{\lambda}\sum\limits_{n=1}^{\infty}\,\frac{Z_{n}^{\text{Disk}}}{(kZ_{1}^{\text{Disk}})^{n}}\,R(\lambda)^{n-1}R_{ij}(\lambda), (7)

where δij/λ\delta_{ij}/\lambda is like “bare propagator” and ZnDiskZ_{n}^{\text{Disk}} is the bulk partition function on a disk topology with nn asymptotic boundaries represented by the black solid arrows and nn blue curved lines for the EOW branes. In a microcanonical ensemble with fixed energy EE, the ratio of the bulk partition functions is simplified as

ZnDisk(Z1Disk)n=e(n1)𝕊,e𝕊:=eS0ρDisk(E)ΔE,ρDisk(E):=sinh(2π2E)2π2,\displaystyle\frac{Z_{n}^{\text{Disk}}}{(Z_{1}^{\text{Disk}})^{n}}=e^{-(n-1)\mathbb{S}},\;\;\;e^{\mathbb{S}}:=e^{S_{0}}\rho_{\text{Disk}}(E)\Delta E,\;\;\;\rho_{\text{Disk}}(E):=\frac{\sinh(2\pi\sqrt{2E})}{2\pi^{2}}, (8)

where EE dependence appears through 𝕊\mathbb{S}, and ΔE\Delta E is the width of the microcanonical energy window. Performing the infinite sum in eq. (7), we obtain

R(λ)2+(eSkλkeS)R(λ)+k2eSλ=0,\displaystyle R(\lambda)^{2}+\left(\,\frac{e^{\textbf{S}}-k}{\lambda}-ke^{\textbf{S}}\,\right)\,R(\lambda)+\dfrac{k^{2}e^{\textbf{S}}}{\lambda}\,=0, (9)

and a solution of R(λ)R(\lambda) with the asymptotic behavior R(λ)k/λR(\lambda)\to k/\lambda at λ+\lambda\to+\infty is

R(λ)\displaystyle R(\lambda) =keS2λ((eSk1)+λ(λλ+)(λλ))(forλ>λ+),\displaystyle=\frac{ke^{\textbf{S}}}{2\lambda}\left(\left(e^{-\textbf{S}}-k^{-1}\right)+\lambda-\sqrt{(\lambda-\lambda_{+})(\lambda-\lambda_{-})}\right){\qquad\left(\mbox{for}\,\,\,\lambda>\lambda_{+}\right)}, (10)
whereλ±:=(k12±eS/2)2.\displaystyle\qquad\qquad\mbox{where}\quad\lambda_{\pm}:=\left(k^{-\frac{1}{2}}\pm e^{-\textbf{S}/2}\right)^{2}. (11)

R(λ)R(\lambda) for λ<λ+\lambda<\lambda_{+} can be obtained by the analytic continuation. From the definition of R(λ)R(\lambda) in (6), using

1λ+iϵ=P(1λ)iπδ(λ),\displaystyle\frac{1}{\lambda+i\epsilon}=\mbox{P}\left(\frac{1}{\lambda}\right)-i\pi\delta(\lambda), (12)

the spectral density D(λ)D(\lambda) of ρ\rho_{\mathbb{R}} is given by

D(λ)\displaystyle D(\lambda) =1πImR(λ+iϵ)\displaystyle=-\frac{1}{\pi}\imaginary R(\lambda+i\epsilon)
=keS2πλ(λλ)(λ+λ)θ(λλ)θ(λ+λ)+(ke𝐒)δ(λ)θ(ke𝐒),\displaystyle=\frac{ke^{\textbf{S}}}{2\pi\lambda}\sqrt{(\lambda-\lambda_{-})(\lambda_{+}-\lambda)}\theta(\lambda-\lambda_{-})\theta(\lambda_{+}-\lambda)+\left(k-e^{\mathbf{S}}\right)\delta(\lambda)\theta(k-e^{\mathbf{S}})\,, (13)

where θ(λ)\theta(\lambda) is the Heaviside step function111From (11), we have R(λ)=keS2λ((eSk1)+λ+(λ+λ)(λλ))for(λ+λλ0).\displaystyle R(\lambda)=\frac{ke^{\textbf{S}}}{2\lambda}\left(\left(e^{-\textbf{S}}-k^{-1}\right)+\lambda+\sqrt{(\lambda_{+}-\lambda)(\lambda_{-}-\lambda)}\right){\qquad\mbox{for}\quad\left(\lambda_{+}\geq\lambda_{-}\geq\lambda\geq 0\right)}. (14) The relative sign in front of square root changes between λ>λ+\lambda>\lambda_{+} and 0λ<λ0\leq\lambda<\lambda_{-} because we change both of the argument θ+\theta_{+} and θ\theta_{-} by π\pi in λλ+:=r+eiθ+\lambda-\lambda_{+}:=r_{+}e^{i\theta_{+}} and λλ:=reiθ\lambda-\lambda_{-}:=r_{-}e^{i\theta_{-}}. See, for instance, Lu:2014jua . Thus, λ=0\lambda=0 pole in R(λ)R(\lambda) gives a Dirac delta function proportional to keS2(eSk1+λ+λ)=(keS)θ(keS).\displaystyle\frac{{ke^{\textbf{S}}}}{2}\left(e^{-\textbf{S}}-k^{-1}+\sqrt{\lambda_{+}\lambda_{-}}\right)=\left(k-e^{\textbf{S}}\right)\theta(k-e^{\textbf{S}}). (15) .

One can check that the normalization of D(λ)D(\lambda) is

D(λ)𝑑λ=k,D(λ)λ𝑑λ=1.\displaystyle\int D(\lambda)d\lambda=k,\quad\int D(\lambda)\lambda\,d\lambda=1. (16)

The first normalization means that the size of ρ\rho_{\mathbb{R}} is kk, and the second normalization means that Trρ=1\Tr_{\mathbb{R}}\rho_{\mathbb{R}}=1. D(λ)D(\lambda) is simplified as

D(λ)=k22πλλ(4kλ)θ(λ)θ(4kλ),whenk=eS.\displaystyle D(\lambda)=\frac{k^{2}}{2\pi\lambda}\sqrt{\lambda\left(\frac{4}{k}-\lambda\right)}\theta(\lambda)\theta\left(\frac{4}{k}-\lambda\right)\,,\quad\mbox{when}\quad k=e^{\textbf{S}}\,. (17)

Using (11) and (13), the entanglement entropy SS_{\mathbb{R}} of the auxiliary system \mathbb{R} can be calculated as

S=\displaystyle S_{\mathbb{R}}= 𝑑λD(λ)λlogλ\displaystyle\;-\int d\lambda D(\lambda)\lambda\log\lambda
=\displaystyle= keS2πλλ+𝑑λ(λλ)(λ+λ)logλ.\displaystyle\;-\frac{ke^{\textbf{S}}}{2\pi}\int_{\lambda_{-}}^{\lambda_{+}}d\lambda\sqrt{(\lambda-\lambda_{-})(\lambda_{+}-\lambda)}\log\lambda. (18)

This integral can be computed exactly, and the result is

S=logmm2n,m:=min{k,eS},n:=max{k,eS},\displaystyle S_{\mathbb{R}}=\log m-\frac{m}{2n},\;\;\;m:=\min\{k,e^{\textbf{S}}\},\;\;\;n:=\max\{k,e^{\textbf{S}}\}, (19)

which perfectly matches the Page’s result for nm1n\geq m\gg 1 Page:1993df . If k=eSk=e^{\textbf{S}}, the entanglement entropy is

S=logk12,ifk=eS.\displaystyle S_{\mathbb{R}}=\log k-\frac{1}{2}\,,\quad\mbox{if}\quad k=e^{\textbf{S}}. (20)

2.2 Review of the IOP matrix model

The IOP matrix model Iizuka:2008eb is a matrix model given by the following Hamiltonian

HIOP=mAijAji+Maiai+Hint,Hint=haialAijAjl,\displaystyle H_{IOP}=mA_{ij}^{\dagger}A_{ji}+Ma_{i}^{\dagger}a_{i}+H_{int},\;\;\;H_{int}=ha_{i}^{\dagger}a_{l}A_{ij}^{\dagger}A_{jl}, (21)

where the sum of subscripts is taken from 11 to NN. Here, aia_{i} is the annihilation operator for a harmonic oscillator in the fundamental of U(N)U(N), and AijA_{ij} is the annihilation operator for an oscillator in the adjoint. This matrix model was introduced as a toy model of the gauge dual of an AdS black hole, where the adjoint fields can be interpreted as background NN D0-branes for the black hole, and the fundamental fields can be interpreted as strings stretched from a probe D0-brane.

To solve the spectrum density analytically in the large NN limit with fixed ’t Hooft coupling λ’t Hooft:=hN\lambda_{\text{'t Hooft}}:=hN, we also take the large MM limit MmM\gg m and MTM\gg T so that aiai0a_{i}^{\dagger}a_{i}\sim 0 in the thermal ensemble at finite temperature TT. We consider the following time-ordered Green’s functions at finite temperature

eiMtTai(t)aj(0)T\displaystyle e^{iMt}\Big{\langle}\mbox{T}\,a_{i}(t)\,a_{j}^{\dagger}(0)\Big{\rangle}_{T} =:G(t)δij,\displaystyle=:G(t)\delta_{ij}, (22)
TAij(t)Akl(0)T\displaystyle\Big{\langle}\mbox{T}A_{ij}(t)A_{kl}^{\dagger}(0)\Big{\rangle}_{T} =:L(t)δilδjk.\displaystyle=:L(t)\delta_{il}\delta_{jk}. (23)

With the Fourier transformation f~(ω)=𝑑teiωtf(t)\tilde{f}(\omega)=\int dt\,e^{i\omega t}f(t), free thermal propagators in frequency space are given by

G~0(ω)\displaystyle\tilde{G}_{0}(\omega) =iω+iϵ,\displaystyle=\frac{i}{\omega+i\epsilon}, (24)
L~0(ω)\displaystyle\tilde{L}_{0}(\omega) =i1y(1ωm+iϵyωmiϵ),y:=em/T,\displaystyle=\frac{i}{1-y}\left(\frac{1}{\omega-m+i\epsilon}-\frac{y}{\omega-m-i\epsilon}\right),\;\;\;y:=e^{-m/T}, (25)

where G~0(ω)\tilde{G}_{0}(\omega) does not depend on TT in the large MM limit, and L~(ω)\tilde{L}(\omega) in the large NN limit becomes the free propagator L~0(ω)\tilde{L}_{0}(\omega) since the backreaction from the fundamental is suppressed by 1/N1/N.

Refer to caption
Figure 2: Schwinger-Dyson equation for the IOP matrix model in the planar limit.

In the limit where NN and MM are large, the Schwinger-Dyson equation of G~(ω)\tilde{G}(\omega) is shown in Fig. 2, which has the same graphical structure as the Schwinger-Dyson equation of R(λ)R(\lambda) in the PSSY model. See Figure 2 of Iizuka:2008eb and Figure (2.25) of Penington:2019kki as well. The Schwinger-Dyson equation of G~(ω)\tilde{G}(\omega) is given by

G~(ω)=G~0(ω)+yG~0(ω)G~(ω)n=0(iλ’t Hooft1y)n+1G~(ω)n.\displaystyle\tilde{G}(\omega)=\tilde{G}_{0}(\omega)+y\tilde{G}_{0}(\omega)\tilde{G}(\omega)\sum_{n=0}^{\infty}\left(\frac{-i\lambda_{\text{'t Hooft}}}{1-y}\right)^{n+1}\tilde{G}(\omega)^{n}. (26)

Performing the infinite sum, we obtain

G~(ω)2i(1y)(1ω+1λ’t Hooft)G~(ω)1yωλ’t Hooft=0,\displaystyle\tilde{G}(\omega)^{2}-i(1-y)\left(\frac{1}{\omega}+\frac{1}{\lambda_{\text{'t Hooft}}}\right)\tilde{G}(\omega)-\frac{1-y}{\omega\lambda_{\text{'t Hooft}}}=0, (27)

and its solution is

G~(ω)\displaystyle\tilde{G}(\omega) =i2ω1yλ’t Hooft(λ’t Hooft1y(1y)+ω(ωω+)(ωω))(forω>ω+),\displaystyle=\frac{i}{2\omega}\frac{1-y}{\lambda_{\text{'t Hooft}}}\left(\frac{\lambda_{\text{'t Hooft}}}{1-y}(1-y)+\omega-\sqrt{\left(\omega-\omega_{+}\right)\left(\omega-\omega_{-}\right)}\right)\,\,\,(\mbox{for}\,\,\,\omega>\omega_{+})\,, (28)
whereω±:=λ’t Hooft1y(1±y)20,\displaystyle\qquad\qquad\mbox{where}\quad\omega_{\pm}:=\frac{\lambda_{\text{'t Hooft}}}{1-y}\left(1\pm\sqrt{y}\right)^{2}\geq 0\,, (29)

where 0y10\leq y\leq 1 and we take the branch such that G~(ω)\tilde{G}(\omega) at ω+\omega\to+\infty becomes the free propagator given by (24). G~(ω)\tilde{G}(\omega) for ω<ω+\omega<\omega_{+} can be obtained by the analytic continuation. Again using (12) and (28), the spectral density F(ω)F(\omega) of the two-point function of fundamental fields is obtained as222From (28), we have G~(ω)\displaystyle\tilde{G}(\omega) =i2ω1yλ’t Hooft(λ’t Hooft1y(1y)+ω+(ω+ω)(ωω)),(forω+ωω>0).\displaystyle=\frac{i}{2\omega}\frac{1-y}{\lambda_{\text{'t Hooft}}}\left(\frac{\lambda_{\text{'t Hooft}}}{1-y}(1-y)+\omega+\sqrt{\left(\omega_{+}-\omega\right)\left(\omega_{-}-\omega\right)}\right),\,\,\,(\mbox{for}\,\,\,\omega_{+}\geq\omega_{-}\geq\omega>0)\,. (30) Again, the relative sign in front of the square root changes between ω>ω+\omega>\omega_{+} and 0ω<ω0\leq\omega<\omega_{-}. Thus, ω0\omega\to 0 pole gives a Dirac delta function proportional to 121yλ’t Hooft(λ’t Hooft1y(1y)+ω+ω)=(1y)θ(1y).\displaystyle\frac{1}{2}\frac{1-y}{\lambda_{\text{'t Hooft}}}\left(\frac{\lambda_{\text{'t Hooft}}}{1-y}(1-y)+\sqrt{\omega_{+}\omega_{-}}\right)=(1-y)\theta(1-y). (31)

F(ω)\displaystyle F(\omega) =1πReG~(ω+iϵ)\displaystyle=\frac{1}{\pi}\real\tilde{G}(\omega+i\epsilon) (32)
=12πω1yλ’t Hooft(ωω)(ω+ω)θ(ωω)θ(ω+ω)\displaystyle=\frac{1}{2\pi\omega}\frac{1-y}{\lambda_{\text{'t Hooft}}}\sqrt{(\omega-\omega_{-})(\omega_{+}-\omega)}\theta(\omega-\omega_{-})\theta(\omega_{+}-\omega)
+(1y)θ(1y)δ(ω).\displaystyle\qquad+(1-y)\theta(1-y)\delta(\omega)\,. (33)

Note that our convention of the propagators includes a factor ii in the numerator as seen in eq. (24). F(ω)F(\omega) is normalized as

F(ω)𝑑ω=1,F(ω)ω𝑑ω=yλ’t Hooft1y.\displaystyle\int F(\omega)d\omega=1,\;\;\;\int F(\omega)\omega d\omega=\frac{y\lambda_{\text{'t Hooft}}}{1-y}. (34)

3 The PSSY model and the IOP matrix model correspondence

3.1 Feynman diagram correspondence between the PSSY model and the IOP matrix model

As seen in Fig. 1 and 2, in the planar limit, the Schwinger-Dyson equations in the PSSY model and the IOP matrix model have the same graphical structure. From now on, we elaborate on the correspondence of each diagram.

From diagrams in the IOP matrix model, one can uniquely construct the corresponding diagrams in the PSSY model and vice versa. The Feynman diagram correspondence can be obtained by the following prescription. See Fig. 3.

  1. 1.

    Extend vertices in the IOP matrix model horizontally and draw straight lines with horizontal arrows from right to left. These arrows represent the asymptotic boundaries with the time direction from the ket to the bra in the PSSY model.

  2. 2.

    Rewrite the adjoint correlators in the IOP matrix model as blue solid curves in the PSSY model. These blue solid curves correspond to EOW branes in the PSSY model.

  3. 3.

    Fill in regions above the right-to-left horizontal arrows corresponding to asymptotic boundaries with a gray shadow. These shaded regions correspond to bulk geometries in the PSSY model.

Refer to caption
Figure 3: The prescription to change the IOP vertex (left) to the PSSY vertex (right).

Figure 4 shows examples of corresponding planar diagrams, where we omit arrows in the IOP matrix model for easier comparison. Due to the correspondence between these diagrams, there is also the correspondence between the solutions of the Schwinger-Dyson equations in the planar limit. The correspondence between parameters in both models is examined in the next subsection.

Due to the correspondence, there is one-to-one Feynman diagram correspondence between the IOP matrix model Feynman diagrams and the PSSY model Feynman diagrams. Thus, the correspondence goes beyond the planar limit. For example, Figure 5 shows examples of corresponding non-planar diagrams. From the perspective of the PSSY model, the left figure includes two bulk geometries with a crossing, and the right figure includes a twisted bulk geometry that is anchored to the asymptotic boundaries.

Refer to caption
Figure 4: Corresponding planar diagrams in the IOP matrix model (upper diagrams) and the PSSY model (lower diagrams).
Refer to caption
Figure 5: Corresponding non-planar diagrams in the IOP matrix model (upper diagrams) and the PSSY model (lower diagrams).

Let us look into a little more on the twisted bulk geometry in Figure 5. We can construct this twisted bulk geometry from a bulk geometry for Z3DiskZ_{3}^{\text{Disk}} as follows. First, prepare the bulk geometry for Z3DiskZ_{3}^{\text{Disk}} shown at left in Figure 6. Next, fold the top part downward so that the yellow reverse side is visible as shown in the middle figure. Finally, twist the folded part so that the middle arrow is facing left as shown in the right figure.

Refer to caption
Figure 6: How to construct a twisted bulk geometry (right diagram) by twisting a bulk geometry for Z3DiskZ_{3}^{\text{Disk}} (left diagram). The yellow-shaded surface represents the reverse side of the gray surface.

Following our prescription in reverse, we can also construct the corresponding diagrams in the IOP matrix model from the ones in the PSSY model. However, note that not all bulk geometries in the PSSY model correspond to diagrams in the IOP matrix model. To see this point, for instance, let us consider three examples of bulk geometries that contribute to Tr(ρ6)\Tr(\rho^{6}_{\mathbb{R}}) as shown in Figure 7. In the IOP matrix model, there are diagrams corresponding to the planar one and the non-planar one with a crossing such as the left and middle geometries in Figure 7, respectively. However, there is no diagram in the IOP matrix model for the non-planar geometry with an extra handle such as the right geometry where non-planar effects are due to the extra handle in bulk, not crossings.

Refer to caption
Figure 7: Three examples of bulk geometries that contribute to Tr(ρ6)\Tr(\rho^{6}_{\mathbb{R}}). The left figure is a planar geometry, the middle figure is a non-planar geometry with a crossing, and the right figure is a non-planar geometry with an extra handle in bulk.

These clearly show that the correspondence works as long as we neglect the extra-handle-in-bulk diagrams. Thus in this paper, using the PSSY and the IOP model correspondence, in Subsection 3.3, we calculate the exact non-planar effects associated with a crossing as the middle figure in Fig. 7.

3.2 Parameter correspondence between the PSSY model and the IOP matrix model

Given the one-to-one Feynman diagram correspondence, it is straightforward to read off the parameter correspondence between them. One can observe that the spectral densities D(λ)D(\lambda) in (13) and F(ω)F(\omega) (33) have the same structures. In fact, after rescaling, the spectral density of the IOP matrix model at infinite temperature limit agrees with the spectral density D(λ)D(\lambda) (17) of the PSSY model when k=e𝕊k=e^{\mathbb{S}}.

The reason why we need to take an infinite temperature in the IOP matrix model is as follows. In the propagator L~0(ω)\tilde{L}_{0}(\omega) (25) of the IOP matrix model, there is a difference of a factor yy between the two terms. See eq. (3.1) of Iizuka:2008eb as well. However, there is no such difference in the PSSY model side. To eliminate this difference, we need to take the following infinite temperature limit

y=em/T1andλ’t Hooft0withλy:=λ’t Hooft1y=fixed.\displaystyle y=e^{-m/T}\to 1\quad\text{and}\quad\lambda_{\text{'t Hooft}}\to 0\,\quad\text{with}\;\,\lambda_{y}:=\frac{\lambda_{\text{'t Hooft}}}{1-y}=\mbox{fixed}. (35)

In this infinite temperature limit, the spectral density F(ω)F(\omega) (33) becomes

F(ω)=\displaystyle F(\omega)= 12πωλyω(4λyω)θ(ω)θ(4λyω).\displaystyle\;\frac{1}{2\pi\omega\lambda_{y}}\sqrt{\omega(4\lambda_{y}-\omega)}\theta(\omega)\theta(4\lambda_{y}-\omega). (36)

Since there is a correspondence between the Feynman diagrams in the IOP matrix model and the PSSY model, this F(ω)F(\omega) should correspond to D(λ)D(\lambda) up to some normalization.

Let us compare D(λ)D(\lambda) in the PSSY model given by (13) and F(ω)F(\omega) in the IOP matrix model given by (33). Then it is clear that under the y1y\to 1 limit, one needs k=e𝕊k=e^{\mathbb{S}} limit for the correspondence to work333We will later see in the discussion section that in order to go beyond k=e𝕊k=e^{\mathbb{S}} limit, one needs to consider a rectangular model. As long as we are considering a square matrix in the IOP model, one has to take k=e𝕊k=e^{\mathbb{S}} limit for the correspondence to the PSSY model to work. Note that even in the rectangular model, one always needs y1y\to 1 limit in the IOP model for the correspondence to the PSSY model for ke𝕊k\neq e^{\mathbb{S}}.. Thus we focus on this limit. Furthermore, in order to take into account the normalization difference between F(ω)F(\omega) in the IOP matrix model (34) and D(λ)D(\lambda) in the PSSY model (16), we divide D(λ)D(\lambda) in (17) by kk as

1kD(λ)=k2πλλ(4kλ)θ(λ)θ(4kλ),whenk=eS.\displaystyle\frac{1}{k}D(\lambda)=\frac{k}{2\pi\lambda}\sqrt{\lambda\left(\frac{4}{k}-\lambda\right)}\theta(\lambda)\theta\left(\frac{4}{k}-\lambda\right)\,,\quad\mbox{when}\quad k=e^{\textbf{S}}\,. (37)

Let’s compare (36) and (37). One might naively think that ω\omega in the IOP matrix model corresponds to λ\lambda in the PSSY model. However, this cannot be true since their dimensions do not match. To make ω\omega dimensionless and also to match the parameter range in θ\theta-functions, we define

ω~:=ωλyksuch that0ω4λy 0ω~4k.\displaystyle\tilde{\omega}:=\frac{\omega}{\lambda_{y}k}\quad\mbox{such that}\quad 0\leq\omega\leq 4\lambda_{y}\,\Leftrightarrow\,0\leq\tilde{\omega}\leq\frac{4}{k}. (38)

Then, we can define

F~(ω~):=λykF(ω)such that𝑑ω~F~(ω~)=𝑑ωF(ω)=1.\displaystyle\tilde{F}(\tilde{\omega}):=\lambda_{y}kF(\omega)\quad\mbox{such that}\quad\int d\tilde{\omega}\tilde{F}(\tilde{\omega})=\int d\omega F(\omega)=1. (39)

Thus, we obtain

F~(ω~)=k2πω~ω~(4kω~)θ(ω~)θ(4kω~).\displaystyle\tilde{F}(\tilde{\omega})=\frac{k}{2\pi\tilde{\omega}}\sqrt{\tilde{\omega}\left(\frac{4}{k}-\tilde{\omega}\right)}\theta\left(\tilde{\omega}\right)\theta\left(\frac{4}{k}-\tilde{\omega}\right). (40)

It is then clear that there is a parameter correspondence between the two models as follows

ω~(IOP)λ(PSSY),F~(ω~)(IOP)1kD(λ)(PSSY),\displaystyle\tilde{\omega}\,(\mbox{IOP})\,\leftrightarrow\,\lambda\,(\mbox{PSSY})\,,\quad\tilde{F}(\tilde{\omega})\,(\mbox{IOP})\,\leftrightarrow\,\frac{1}{k}D(\lambda)\,(\mbox{PSSY})\,, (41)
N(IOP)k=e𝕊(PSSY)\displaystyle\qquad\qquad\qquad N\,(\mbox{IOP})\,\leftrightarrow\,k=e^{\mathbb{S}}\,(\mbox{PSSY})\, (42)

at y1y\to 1 limit. Note that, for the planar limit, we consider the large NN limit in the IOP matrix model and the large kk, e𝕊e^{\mathbb{S}} limit in the PSSY model, and they correspond as (42).

Let us investigate the correspondence in more detail. In the PSSY model, the spectral density D(λ)D(\lambda) is computed from the resolvent R(λ)R(\lambda)

R(λ)=Tr1λ𝟙ρ=l=1kl|1λ𝟙ρ|l,\displaystyle R(\lambda)=\Tr\frac{1}{\lambda\mathds{1}-\rho_{\mathbb{R}}}=\sum_{l=1}^{k}\bra{l}_{\mathbb{R}}\frac{1}{\lambda\mathds{1}-\rho_{\mathbb{R}}}\ket{l}_{\mathbb{R}}, (43)
whereρ=1ki,j=1k|ji|ψi|ψj𝔹.\displaystyle\mbox{where}\quad\rho_{\mathbb{R}}=\frac{1}{k}\,\sum\limits_{i,j=1}^{k}\,\ket{j}\bra{i}_{\mathbb{R}}\,\braket{\psi_{i}}{\psi_{j}}_{\mathbb{B}}. (44)

In the IOP matrix model, the two-point function G~(ω)\tilde{G}(\omega) in the large MM limit can be expressed as

NG~(ω)=l=1Naliω𝟙HintalT\displaystyle N\tilde{G}(\omega)=\sum_{l=1}^{N}\Big{\langle}a_{l}\frac{i}{\omega\mathds{1}-H_{int}}a^{\dagger}_{l}\Big{\rangle}_{T}
iλyN2G~(ω)=l=1Nal1ωλyN𝟙1yN2ajaiAjkAkialT.\displaystyle\implies\quad-i\lambda_{y}N^{2}\tilde{G}(\omega)=\sum_{l=1}^{N}\Big{\langle}a_{l}\frac{1}{\frac{\omega}{\lambda_{y}N}\mathds{1}-\frac{1-y}{N^{2}}a_{j}^{\dagger}a_{i}A^{\dagger}_{jk}A_{ki}}a^{\dagger}_{l}\Big{\rangle}_{T}. (45)

Since we take MM to be large so that the number of fundamental fields is always one in the evaluation of G~(ω)\tilde{G}(\omega), we can treat al|va_{l}^{\dagger}\ket{v} as an NN-dimensional one-particle excited state basis, where |v\ket{v} is the ground state for the fundamental field. Comparing eqs.  (44) and (45), we obtain the following additional relationships444To be precise, since the trace of a matrix is invariant under the transformation of a basis, there is the ambiguity of a unitary matrix UU in the correspondence (47) as HintλyN(IOP)\displaystyle\qquad\frac{H_{int}}{\lambda_{y}N}\,\,\,\,(\mbox{IOP})\, UρU(PSSY).\displaystyle\leftrightarrow\,\,\,\,U\rho_{\mathbb{R}}U^{\dagger}\,\,\,\,(\mbox{PSSY})\,. (46)

HintλyN=1yN2ajaiAjkAki(IOP)\displaystyle\qquad\frac{H_{int}}{\lambda_{y}N}=\frac{1-y}{N^{2}}a_{j}^{\dagger}a_{i}A^{\dagger}_{jk}A_{ki}\,(\mbox{IOP})\, ρ(PSSY),\displaystyle\leftrightarrow\,\,\,\,\rho_{\mathbb{R}}\,\,\,\,(\mbox{PSSY})\,, (47)
iλyN2G~(ω)(IOP)\displaystyle-i\lambda_{y}N^{2}\tilde{G}(\omega)\,\,\,(\mbox{IOP})\, R(λ)(PSSY).\displaystyle\leftrightarrow\,\,R(\lambda)\,(\mbox{PSSY})\,. (48)

in addition to the parameter correspondence given by (41) and (42). In (47), naively one might wonder if this term vanishes in the y1y\to 1 limit. However, the adjoint propagator is also proportional to 1/(1y)1/(1-y) as seen in (25), thus this is a well-defined limit even in y1y\to 1.

Furthermore, |l\ket{l}_{\mathbb{R}}, which forms an orthonormal basis for the radiation Hilbert space, corresponds to the one-fundamental excited state al|va_{l}^{\dagger}\ket{v} that is again orthogonal. Given these correspondences, one can also see the relationship

Random ensemble average of ψi|ψj𝔹(PSSY)\displaystyle\text{Random ensemble average of }\braket{\psi_{i}}{\psi_{j}}_{\mathbb{B}}\,(\mbox{PSSY})
 Expectation value of 1yNAjlAli(IOP).\displaystyle\leftrightarrow\,\,\text{ Expectation value of }\frac{1-y}{N}A^{\dagger}_{jl}A_{li}\,\,\,(\mbox{IOP}). (49)

In the PSSY model, the Gaussian random property of ψi|ψj𝔹\braket{\psi_{i}}{\psi_{j}}_{\mathbb{B}} is crucial for connected wormhole contributions. From the viewpoint of the IOP matrix model, this Gaussian randomness comes from the fact that the adjoint fields AA^{\dagger} behave like Gaussian free fields. In random matrix theory, the spectral density D(λ)D(\lambda) (13) up to the normalization is known as the Marchenko-Pastur distribution Pastur:1967zca , see, for instance, Muck:2024fpb . The reason why the Marchenko-Pastur distribution appears is that ψi|ψj𝔹\braket{\psi_{i}}{\psi_{j}}_{\mathbb{B}} in the PSSY model can be interpreted as a Gram matrix Balasubramanian:2022gmo ; Balasubramanian:2022lnw ; Climent:2024trz and HintH_{int} in the IOP matrix model is proportional to AAA^{\dagger}A.

3.3 Non-planar correction of the entanglement entropy in the PSSY model via the IOP matrix model correspondence

Non-planar 1/N21/N^{2} correction of the two-point function G~(ω)\tilde{G}(\omega) in the IOP matrix model was computed by Iizuka:2008eb . By using the PSSY model and the IOP matrix model correspondence, it is straightforward to obtain the non-planar 1/k21/k^{2} correction of the reduced density matrix and its von Neumann entropy in the PSSY model. Especially, the spectral density D(λ)D(\lambda) (17) in the PSSY model and the rescaled one F~(ω~)\tilde{F}(\tilde{\omega}) (40) in the IOP matrix model corresponds in the planar limit. Then the non-planar 1/N21/N^{2} correction of the von Neumann entropy in the IOP matrix model would be a part of the non-planar 1/k21/k^{2} correction of the entanglement entropy in the PSSY model.

The non-planar 1/N21/N^{2} correction of G~(ω)\tilde{G}(\omega) is calculated in Iizuka:2008eb , and it is

G~(ω)\displaystyle\tilde{G}(\omega) =G~(0)(ω)+1N2G~(1)(ω)+𝒪(1N4),\displaystyle=\tilde{G}^{(0)}(\omega)+{1\over N^{2}}\tilde{G}^{(1)}(\omega)+{\cal{O}}\left({1\over N^{4}}\right)\,, (50)
G~(0)(ω)\displaystyle\tilde{G}^{(0)}(\omega) =i2λy(114λyω),\displaystyle=\frac{i}{2\lambda_{y}}\left(1-\sqrt{1-\frac{4\lambda_{y}}{\omega}}\right)\,, (51)
x0\displaystyle x_{0} :=iλyG~(0)(ω)=12(114λyω),\displaystyle:=-i\lambda_{y}\tilde{G}^{(0)}(\omega)=\frac{1}{2}\left(1-\sqrt{1-\frac{4\lambda_{y}}{\omega}}\right)\,, (52)
G~(1)(ω)\displaystyle\tilde{G}^{(1)}(\omega) =ix03(1x0)4(12x0)4(ω(1x0)2λy)=iλy3(ω4λy)5/2ω3/2.\displaystyle=\frac{ix_{0}^{3}(1-x_{0})^{4}}{(1-2x_{0})^{4}(\omega(1-x_{0})^{2}-\lambda_{y})}=\frac{i\lambda_{y}^{3}}{(\omega-4\lambda_{y})^{5/2}\omega^{3/2}}\,. (53)

By using this result, we obtain the non-planar 1/N21/N^{2} correction of the spectral density

F(ω)\displaystyle F(\omega) =F(0)(ω)+1N2F(1)(ω)+𝒪(1N4),\displaystyle=F^{(0)}(\omega)+{1\over N^{2}}F^{(1)}(\omega)+{\cal{O}}\left({1\over N^{4}}\right), (54)
F(0)(ω)\displaystyle F^{(0)}(\omega) =1πReG~(0)(ω)=ω(4λyω)2πλyωθ(ω)θ(4λyω),\displaystyle=\frac{1}{\pi}\real\tilde{G}^{(0)}(\omega)=\frac{\sqrt{\omega(4\lambda_{y}-\omega)}}{2\pi\lambda_{y}\omega}\theta(\omega)\theta\left(4\lambda_{y}-\omega\right), (55)
F(1)(ω)\displaystyle F^{(1)}(\omega) =1πReG~(1)(ω)=λy3πω3/2(4λyω)5/2θ(ω)θ(4λyω).\displaystyle=\frac{1}{\pi}\real\tilde{G}^{(1)}(\omega)=\frac{\lambda_{y}^{3}}{\pi\omega^{3/2}(4\lambda_{y}-\omega)^{5/2}}\theta(\omega)\theta\left(4\lambda_{y}-\omega\right). (56)

Since G~(1)(ω)\tilde{G}^{(1)}(\omega) (53) is a rational function of x0x_{0} and ω\omega, F(1)(ω)F^{(1)}(\omega) has branch points at ω=0,4λy\omega=0,4\lambda_{y} that are the same branch points of F(0)(ω)F^{(0)}(\omega). This property comes from the fact that the perturbation equation determining G~(1)(ω)\tilde{G}^{(1)}(\omega) is written in terms of G~(0)(ω)\tilde{G}^{(0)}(\omega). Note that even though the branch points of F(0)(ω)F^{(0)}(\omega) and F(1)(ω)F^{(1)}(\omega) are the same, F(1)(ω)F^{(1)}(\omega) is singular than F(0)(ω)F^{(0)}(\omega).

Given the correspondence we discussed in the previous subsection, we can read off the 1/k21/k^{2} corrections in the PSSY model from (38), (39), (41), (42), and (56) as

1kD(0)(λ)\displaystyle\frac{1}{k}D^{(0)}(\lambda) =k2πλλ(4kλ)θ(λ)θ(4kλ),\displaystyle=\frac{k}{2\pi\lambda}\sqrt{\lambda\left(\frac{4}{k}-\lambda\right)}\theta(\lambda)\theta\left(\frac{4}{k}-\lambda\right)\,, (57)
1kD(1)(λ)\displaystyle\frac{1}{k}D^{(1)}(\lambda) =1πk31λ3/2(4kλ)5/2θ(λ)θ(4kλ),\displaystyle=\frac{1}{\pi k^{3}}\frac{1}{\lambda^{3/2}(\frac{4}{k}-\lambda)^{5/2}}\theta(\lambda)\theta\left(\frac{4}{k}-\lambda\right)\,, (58)

when k=eSk=e^{\textbf{S}}. Here D(0)(λ)D^{(0)}(\lambda) and D(1)(λ)D^{(1)}(\lambda) are the same order since

1kD(0)(λ)=(k),1kD(1)(λ)=(k)for λ1k.\displaystyle\frac{1}{k}D^{(0)}(\lambda)=\order{k}\,,\quad\frac{1}{k}D^{(1)}(\lambda)=\order{k}\,\quad\mbox{for $\lambda\sim\frac{1}{k}$}. (59)

With this, one can calculate the entanglement entropy for the radiation SS_{\mathbb{R}} as

S\displaystyle S_{\mathbb{R}} :=D(λ)λlogλ\displaystyle:=-\int D(\lambda)\lambda\log\lambda
=(D(0)(λ)+1k2D(1)(λ)+(1k4))λlogλ.\displaystyle=-\int\left(D^{(0)}(\lambda)+\frac{1}{k^{2}}D^{(1)}(\lambda)+\order{1\over k^{4}}\right)\lambda\log\lambda. (60)

The leading term can be evaluated as

D(0)(λ)λlogλ=k22π04/kλ(4kλ)logλdλ=logk12,\displaystyle-\int D^{(0)}(\lambda)\lambda\log\lambda=-\frac{k^{2}}{2\pi}\int_{0}^{4/k}\sqrt{\lambda\left({4\over k}-\lambda\right)}\log\lambda\,d\lambda=\log k-\frac{1}{2}, (61)

which agrees with eq. (20). The subleading term is

1k2D(1)(λ)λlogλdλ\displaystyle-\frac{1}{k^{2}}\int D^{(1)}(\lambda)\lambda\log\lambda\,d\lambda =1πk404/kλλ3/2(4kλ)5/2logλdλ=Ck2,\displaystyle=-\frac{1}{\pi k^{4}}\int_{0}^{4/k}\frac{\lambda}{\lambda^{3/2}\left(\frac{4}{k}-\lambda\right)^{5/2}}\log\lambda\,d\lambda=\frac{C}{k^{2}}, (62)
whereC\displaystyle\mbox{where}\quad C :=1π04x2(x(4x))5/2log(xk)𝑑x,\displaystyle:=-\frac{1}{\pi}\int_{0}^{4}\frac{x^{2}}{\left(x(4-x)\right)^{5/2}}\log\left(\frac{x}{k}\right)dx, (63)

where we change the integration variable λ=x/k\lambda=x/k.

CC does not converge due to more singular nature of D(1)(λ)D^{(1)}(\lambda) than D(0)(λ)D^{(0)}(\lambda). To regularize this integral, we introduce a small cutoff ϵ\epsilon so that CC is regularized as

Cϵ\displaystyle C_{\epsilon} :=1π04ϵx2(x(4x))5/2log(xk)𝑑x\displaystyle:=-\frac{1}{\pi}\int_{0}^{4-\epsilon}\frac{x^{2}}{\left(x(4-x)\right)^{5/2}}\log\left(\frac{x}{k}\right)dx (64)
=logk43πϵ3/2+logk4+28πϵ+112+(ϵ1/2).\displaystyle=\frac{\log\frac{k}{4}}{3\pi\epsilon^{3/2}}+\frac{\log\frac{k}{4}+2}{8\pi\sqrt{\epsilon}}+\frac{1}{12}+\order{\epsilon^{1/2}}. (65)

The first and second terms are divergent but they are regulator dependent. On the other hand, 1/121/12 is a regulator-independent one. Thus, we focus on this 1/121/12 by subtracting the UV divergent terms555One might be able to justify this argument along the line of “renormalized entanglement entropy” of Liu:2012eea ..

Therefore, after subtracting the regulator-dependent divergent terms at ϵ0\epsilon\to 0, we obtain a finite result

S=logk12+112k2+𝒪(1k4),whenk=e𝕊.\displaystyle S_{\mathbb{R}}=\log k-\frac{1}{2}+\frac{1}{12k^{2}}+{\cal{O}}\left({1\over k^{4}}\right)\,,\quad\mbox{when}\quad k=e^{\mathbb{S}}. (66)

Since the leading term (61) in SS_{\mathbb{R}} agrees with SS_{\mathbb{R}} (20) in the planar limit, we expect that the sub-leading term in SS_{\mathbb{R}} (66) corresponds to a part of non-planar 1/k21/k^{2} corrections of SS_{\mathbb{R}} in the PSSY model. As shown in Fig. 7, non-planar corrections of SS_{\mathbb{R}} in the PSSY model come from non-planar diagrams with extra-handle-in-bulk and crossings. We expect that the sub-leading term in SS_{\mathbb{R}} (66) corresponds to the non-planar correction of SS_{\mathbb{R}} from crossings, not extra-handle-in-bulk.

4 Short conclusion and discussions

As mentioned in Subsection 2.1, the entanglement entropy in the PSSY model and the one of a random pure state coincide with each other in the planar limit for large Hilbert space dimensions. We expect that non-planar corrections in the PSSY model and the IOP matrix model have some connection to Page’s conjecture Page:1993df ; Foong:1994eja ; PhysRevE.52.5653 ; Sen:1996ph on the entanglement entropy of a random pure state for general Hilbert space dimensions.

Page’s conjecture on the entanglement entropy SRS_{R} of a random pure state Page:1993df , which was probed by Foong:1994eja ; PhysRevE.52.5653 ; Sen:1996ph , is

SR=k=n+1mn1km12n,\displaystyle S_{R}=\sum_{k=n+1}^{mn}\frac{1}{k}-\frac{m-1}{2n}, (67)

where mm and nn are Hilbert space dimensions of two subsystems, and we assume that nmn\geq m. By expanding SRS_{R} with large nn, we obtain

SR=logm+1m22mn+112112m2n2+𝒪(1n4).\displaystyle S_{R}=\log m+\frac{\frac{1-m^{2}}{2m}}{n}+\frac{\frac{1}{12}-\frac{1}{12m^{2}}}{n^{2}}+{\cal{O}}\left({1\over n^{4}}\right). (68)

When m=n=k=eSm=n=k=e^{\textbf{S}}, this expansion becomes

SR=logk12+(12+112)1k2+𝒪(1k4).\displaystyle S_{R}=\log k-\frac{1}{2}+\left(\frac{1}{2}+\frac{1}{12}\right)\frac{1}{k^{2}}+{\cal{O}}\left({1\over k^{4}}\right). (69)

Let us compare with the non-planar correction of entanglement entropy in the PSSY model computed in eq. (66). In eq. (66), we have 1/12k21/12k^{2}, which appears in (69). This gives a natural prediction that the resummation of the extra-handle-in-bulk diagrams, as shown in the right figure in Fig. 7, should yield 1/2k21/2k^{2}.

How to show 1/2k21/2k^{2} by re-summing all the extra-handle-in-bulk diagrams by an explicit calculation is an open question. The main difficulty is associated with the systematic resummation of all diagrams. Note that each diagram can be explicitly calculated at least in a canonical ensemble using the Weil-Petersson volume as shown in Saad:2019lba . However, if we restrict only some subsets of diagrams and assume that the others do not contribute, one can handle the resummation. For example, one may consider the following ansatz for the Schwinger-Dyson equation

λR(λ)=k+n=1ZnR(λ)nknZ1n,\displaystyle\lambda R(\lambda)=k+\sum\limits_{n=1}^{\infty}Z_{n}\frac{R(\lambda)^{n}}{k^{n}Z_{1}^{n}}, (70)

with

R(λ)=R0(λ)+R1(λ),Zn=ZnDisk(1+a),ZnDisk(Z1Disk)n=eS(n1),\displaystyle R(\lambda)=R^{0}(\lambda)+R^{1}(\lambda),\;\;\;Z_{n}=Z^{\text{Disk}}_{n}(1+a),\;\;\;\frac{Z_{n}^{\text{Disk}}}{(Z_{1}^{\text{Disk}})^{n}}=e^{-\textbf{S}(n-1)}, (71)

where R0(λ)R^{0}(\lambda) is the resolvent in the planar limit (10). We introduce the subleading terms R1(λ)R^{1}(\lambda) and aa, where aa does not depend on λ\lambda and it captures the effects of extra-handle-in-bulk on a disk. Then, we can solve R1(λ)R^{1}(\lambda) perturbatively as a function of aa, which depends on EE in the microcanonical ensemble. Since the black hole entropy S depends on EE, EE is a function of the dimension of the Hilbert space of the subsystem, and thus, EE dependence can be converted into eSe^{\textbf{S}} dependence. See Appendix A for more detail.

Of course, this approach enables us to resum only the subsets of all the diagrams with an extra handle, since the ansatz (71) includes only a disk geometry with an extra handle on it. It does not include two disks connected by a handle such as the double trumpet geometry. For example, the diagram as Fig. 8 is missing. We leave this resummation problem of all the extra-handle-in-bulk diagrams as a future problem.

Refer to caption
Figure 8: An example of bulk geometries with a handle connecting two shaded regions.

So far we have considered the correspondence in the case of k=e𝕊k=e^{\mathbb{S}}. To generalize it to the case of ke𝕊k\neq e^{\mathbb{S}} with y=1y=1, we can consider a rectangular model such that AA^{\dagger} is a rectangular N×KN\times K matrix. The two-point function G~(ω)\tilde{G}(\omega) of the rectangular model in the large NN limit with fixed K/NK/N was derived by Iizuka:2008eb

G~(ω)=\displaystyle\tilde{G}(\omega)= i2ωλ’t Hooft[h(NKy)+ω(1y)(1y)(ωω+)(ωω)],\displaystyle\,\frac{i}{2\omega\lambda_{\text{'t Hooft}}}\left[h(N-Ky)+\omega(1-y)-(1-y)\sqrt{(\omega-\omega_{+})(\omega-\omega_{-})}\right], (72)
ω±=\displaystyle\omega_{\pm}= h1y(N+Ky±2NKy).\displaystyle\,\frac{h}{1-y}\left(N+Ky\pm 2\sqrt{NKy}\right). (73)

In the infinite temperature limit (35), we obtain

G~(ω)=\displaystyle\tilde{G}(\omega)= i2ωλy(λy(1K/N)+ω(ωω+)(ωω)),\displaystyle\,\frac{i}{2\omega\lambda_{y}}\left(\lambda_{y}\left(1-K/N\right)+\omega-\sqrt{\left(\omega-\omega_{+}\right)\left(\omega-\omega_{-}\right)}\right), (74)
ω±=\displaystyle\omega_{\pm}= λy(1±K/N)2.\displaystyle\,\lambda_{y}\left(1\pm\sqrt{K/N}\right)^{2}. (75)

To compare with the PSSY model, let us define the following rescaled ones

G~~(ω~)\displaystyle\tilde{\tilde{G}}(\tilde{\omega}) :=iλyKG~(ω)=K2ω~((K1N1)+ω~(ω~ω~+)(ω~ω~)),\displaystyle:=-i\lambda_{y}K\tilde{G}(\omega)=\frac{K}{2\tilde{\omega}}\left(\left(K^{-1}-N^{-1}\right)+\tilde{\omega}-\sqrt{\left(\tilde{\omega}-\tilde{\omega}_{+}\right)\left(\tilde{\omega}-\tilde{\omega}_{-}\right)}\right), (76)
ω~\displaystyle\tilde{\omega} :=ωλyK,ω~±:=ω±λyK=(N12±K12)2.\displaystyle:=\frac{\omega}{\lambda_{y}K},\;\;\;\tilde{\omega}_{\pm}:=\frac{\omega_{\pm}}{\lambda_{y}K}=\left(N^{-\frac{1}{2}}\pm K^{-\frac{1}{2}}\right)^{2}. (77)

Comparing them with eqs. (10) and (11), there is a relationship between the parameters as follows

ω~(IOP)λ(PSSY),G~~(ω~)(IOP)1kR(λ)(PSSY),\displaystyle\,\,\tilde{\omega}\,(\mbox{IOP})\,\leftrightarrow\,\lambda\,(\mbox{PSSY})\,,\quad\tilde{\tilde{G}}(\tilde{\omega})\,(\mbox{IOP})\,\leftrightarrow\,\frac{1}{k}R(\lambda)\,(\mbox{PSSY})\,, (78)
N(IOP)k(PSSY),K(IOP)e𝕊(PSSY).\displaystyle N\,(\mbox{IOP})\,\leftrightarrow\,k\,(\mbox{PSSY})\,,\qquad\,\,K\,(\mbox{IOP})\,\leftrightarrow\,e^{\mathbb{S}}\,(\mbox{PSSY})\,. (79)

In the IOP matrix model, there is a parameter yy for finite temperature. However, there is no such parameter in the PSSY model, and thus we consider the infinite temperature limit (35). It is interesting to generalize the PSSY model for the correspondence in the case of y1y\neq 1.

Acknowledgements.
The work of NI was supported in part by JSPS KAKENHI Grant Number 18K03619, MEXT KAKENHI Grant-in-Aid for Transformative Research Areas A “Extreme Universe” No. 21H05184. M.N. was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (RS-2023-00245035).

Appendix A Comments on the partial resum approach in the PSSY model

In the PSSY model, the subleading non-planar corrections of SS_{\mathbb{R}} come from two types of geometries such as the middle and right figures in Fig. 7. We estimate the non-planar correction from the geometry with an extra handle on a disk by using a simple ansatz. Note that our ansatz does not include all geometries with an extra handle. We consider a disk geometry with an extra handle and take the partial resum of only these effects among all non-planar geometries. We do not consider two disks connected by a handle such as the double trumpet geometry and leave its resum and evaluation as future work.

The Schwinger-Dyson equation of R(λ)R(\lambda) in the PSSY model is given by

λR(λ)=k+n=1ZnR(λ)nknZ1n.\displaystyle\lambda R(\lambda)=k+\sum\limits_{n=1}^{\infty}Z_{n}\frac{R(\lambda)^{n}}{k^{n}Z_{1}^{n}}. (80)

Then, we consider the following ansatz

R(λ)=R0(λ)+R1(λ),Zn=ZnDisk(1+a),ZnDisk(Z1Disk)n=eS(n1),\displaystyle R(\lambda)=R^{0}(\lambda)+R^{1}(\lambda),\;\;\;Z_{n}=Z^{\text{Disk}}_{n}(1+a),\;\;\;\frac{Z_{n}^{\text{Disk}}}{(Z_{1}^{\text{Disk}})^{n}}=e^{-\textbf{S}(n-1)}, (81)

where R0(λ)R^{0}(\lambda) is the resolvent in the planar limit (10). We introduce the subleading terms R1(λ)R^{1}(\lambda) and aa, where aa does not depend on λ\lambda. We set k=eSk=e^{\textbf{S}}, and R0(λ)R^{0}(\lambda) becomes

R0(λ)=k22(1(14kλ)).\displaystyle R^{0}(\lambda)=\frac{k^{2}}{2}\left(1-\sqrt{\left(1-\frac{4}{k\lambda}\right)}\right). (82)

By substituting our ansatz (81) into the Schwinger-Dyson equation (80), we obtain the following perturbative equation of R1(λ)R^{1}(\lambda)

λR1(λ)=akR0(λ)k2R0(λ)ak3R0(λ)(k2R0(λ))2+R1(λ)k3(k2R0(λ))2,\displaystyle\lambda R^{1}(\lambda)=a\frac{kR^{0}(\lambda)}{k^{2}-R^{0}(\lambda)}-a\frac{k^{3}R^{0}(\lambda)}{(k^{2}-R^{0}(\lambda))^{2}}+R^{1}(\lambda)\frac{k^{3}}{(k^{2}-R^{0}(\lambda))^{2}}, (83)

where we leave only the first order terms proportional to R1(λ)R^{1}(\lambda) or aa. Its solution is

R1(λ)=ak(R0(λ))2k3(k2R0(λ))2λ=ak(kλ1)2λak214kλ(kλ3)2(kλ4).\displaystyle R^{1}(\lambda)=\frac{ak(R^{0}(\lambda))^{2}}{k^{3}-(k^{2}-R^{0}(\lambda))^{2}\lambda}=\frac{ak(k\lambda-1)}{2\lambda}-\frac{ak^{2}\sqrt{1-\frac{4}{k\lambda}}\left(k\lambda-3\right)}{2(k\lambda-4)}. (84)

The spectral density for R1(λ)R^{1}(\lambda) is given by

D1(λ)\displaystyle D^{1}(\lambda) :=1πImR1(λ+iϵ)=ak24kλ1(kλ3)2π(kλ4)θ(λ)θ(4kλ)ak2δ(λ)\displaystyle:=-\frac{1}{\pi}\imaginary R^{1}(\lambda+i\epsilon)=\frac{ak^{2}\sqrt{\frac{4}{k\lambda}-1}\left(k\lambda-3\right)}{2\pi(k\lambda-4)}\theta(\lambda)\theta\left(\frac{4}{k}-\lambda\right)-\frac{ak}{2}\delta(\lambda)
=(ak22π344kλ1ak22π1414kλ1)θ(λ)θ(4kλ)ak2δ(λ),\displaystyle=\left(\frac{ak^{2}}{2\pi}\frac{3}{4}\sqrt{\frac{4}{k\lambda}-1}-\frac{ak^{2}}{2\pi}\frac{1}{4}\frac{1}{\sqrt{\frac{4}{k\lambda}-1}}\right)\theta(\lambda)\theta\left(\frac{4}{k}-\lambda\right)-\frac{ak}{2}\delta(\lambda), (85)

where the delta function term ak2δ(λ)-\frac{ak}{2}\delta(\lambda) comes from ak2λ-\frac{ak}{2\lambda} in R1(λ)R^{1}(\lambda). Note that the branch points λ=0,4/k\lambda=0,4/k of D1(λ)D^{1}(\lambda) are the same branch points of D(λ)D(\lambda) (17) in the planar limit. As explained in the case of the IOP matrix model, this property seems to come from the perturbative equation of R1(λ)R^{1}(\lambda). One can confirm that

𝑑λD1(λ)=0,𝑑λD1(λ)λ=0.\displaystyle\int d\lambda D^{1}(\lambda)=0,\;\;\;\int d\lambda D^{1}(\lambda)\lambda=0. (86)

Correction of entanglement entropy by this spectral density is

S1:=𝑑λD1(λ)λlog(λ)=a2.\displaystyle S_{\mathbb{R}}^{1}:=-\int d\lambda D^{1}(\lambda)\lambda\log(\lambda)=\frac{a}{2}. (87)

Let us specifically compute the value of aa in JT gravity. First, in a microcanonical ensemble, ZnDisk(E)Z_{n}^{\text{Disk}}(E) is given by Penington:2019kki

ZnDisk, microcanonical(E)=eS0ρDisk(E)h(E,μ)nΔE,\displaystyle\qquad\quad Z_{n}^{\text{Disk, microcanonical}}(E)=e^{S_{0}}\rho_{\text{Disk}}(E)h(E,\mu)^{n}\Delta E, (88)
ρDisk(E)=sinh(2π2E)2π2,h(E,μ)=212μ|Γ(μ12+i2E)|2.\displaystyle\rho_{\text{Disk}}(E)=\frac{\sinh\left(2\pi\sqrt{2E}\right)}{2\pi^{2}}\,,\quad h(E,\mu)=2^{1-2\mu}{|\Gamma(\mu-\frac{1}{2}+i\sqrt{2E})|^{2}}. (89)

Next, let us consider the bulk partition function with an extra handle,

Zn1-handle, microcanonical=eS0ρ1-handle(E)h(E,μ)nΔE.\displaystyle Z_{n}^{\text{1-handle, microcanonical}}=e^{-S_{0}}\rho_{\text{1-handle}}(E)h(E,\mu)^{n}\Delta E. (90)

where ρ1-handle(E)\rho_{\text{1-handle}}(E) can be obtained Saad:2019lba using the explicit expression of the Weil-Petersson volume Mirzakhani:2006fta as

ρ1-handle(E)\displaystyle\rho_{\text{1-handle}}(E) =0b𝑑bV1,1(b)ρTrumpet(E,b),\displaystyle=\int_{0}^{\infty}bdbV_{1,1}(b)\rho_{\text{Trumpet}}(E,b)\,, (91)
V1,1(b)\displaystyle V_{1,1}(b) =124(b2+4π2).\displaystyle=\frac{1}{24}(b^{2}+4\pi^{2})\,. (92)

ρTrumpet(E,b)\rho_{\text{Trumpet}}(E,b) is known explicitly Saad:2019pqd as

ρTrumpet(E,b)\displaystyle\rho_{\text{Trumpet}}(E,b) =cos(b2E)π2E.\displaystyle=\frac{\cos(b\sqrt{2E})}{\pi\sqrt{2E}}\,. (93)

However, the integral (91) for bb does not converge. To make it convergent, one can introduce a regulator ebζe^{-b\zeta} in the integrand of (91) as

ρ1-handleζ(E)\displaystyle\rho^{\zeta}_{\text{1-handle}}(E) =0b𝑑bV1,1(b)ρTrumpet(E,b)ebζ=34π2E48πE22E+(ζ2).\displaystyle=\int_{0}^{\infty}bdbV_{1,1}(b)\rho_{\text{Trumpet}}(E,b)e^{-b\zeta}=\frac{3-4\pi^{2}E}{48\pi E^{2}\sqrt{2E}}+\order{\zeta^{2}}. (94)

Thus, in the limit vanishing regulator ζ0\zeta\to 0, one obtain

ρ1-handle(E)=34π2E48πE22E.\displaystyle\rho_{\text{1-handle}}(E)=\frac{3-4\pi^{2}E}{48\pi E^{2}\sqrt{2E}}\,. (95)

By combing (88) and (90), we obtain

ZnDisk, microcanonical+Zn1-handle, microcanonical\displaystyle Z_{n}^{\text{Disk, microcanonical}}+Z_{n}^{\text{1-handle, microcanonical}}
=eS0ρDisk(E)h(E,μ)nΔE(1+ρ1-handle(E)e2S0ρDisk(E)).\displaystyle\qquad=e^{S_{0}}\rho_{\text{Disk}}(E)h(E,\mu)^{n}\Delta E\left(1+\frac{\rho_{\text{1-handle}}(E)}{e^{2S_{0}}\rho_{\text{Disk}}(E)}\right). (96)

Therefore, in JT gravity, aa in our ansatz (81) is given by

a=Zn1-handle, microcanonicalZnDisk, microcanonical=ρ1-handle(E)e2S0ρDisk(E),\displaystyle a=\frac{Z_{n}^{\text{1-handle, microcanonical}}}{Z_{n}^{\text{Disk, microcanonical}}}=\frac{\rho_{\text{1-handle}}(E)}{e^{2S_{0}}\rho_{\text{Disk}}(E)}, (97)

which is a function of the fixed energy EE in the microcanonical ensemble. Moreover, aa is proportional to 1e2S0\frac{1}{e^{2S_{0}}} as expected.

Let us express aa as a function of S. From eq. (8), we obtain E(SS0)28π2E\sim\frac{(\textbf{S}-S_{0})^{2}}{8\pi^{2}} when EE is large. Therefore using (89), (95) and (97), aa can be expressed under this approximation of large EE as

a=ρ1-handle(E)e2S0ρDisk(E)16π6(6(SS0)2)3eS+S0(SS0)5.\displaystyle a=\frac{\rho_{\text{1-handle}}(E)}{e^{2S_{0}}\rho_{\text{Disk}}(E)}\sim\frac{16\pi^{6}(6-(\textbf{S}-S_{0})^{2})}{3e^{\textbf{S}+S_{0}}(\textbf{S}-S_{0})^{5}}. (98)

The expression of aa depends not only on S, which is related to the dimension of Hilbert space of the subsystem, but also on S0S_{0}. This result means that, in contrast to Page’s conjecture (69), aa cannot be expressed by the dimension of Hilbert space only. This discrepancy with Page’s conjecture might be resolved by doing the resum including all geometries with an extra handle.

References