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, Shinshu University,
3-1-1 Asahi, Matsumoto 390-8621, Japan
bbinstitutetext: Institute of Physics, Meiji Gakuin University,
1518 Kamikurata-cho, Totsuka-ku, Yokohama 244-8539, Japan

Page curve from dynamical branes in JT gravity

Kazumi Okuyama b    and Kazuhiro Sakai [email protected], [email protected]
Abstract

We study the Page curve of an evaporating black hole using a toy model given by Jackiw-Teitelboim gravity with Fateev-Zamolodchikov-Zamolodchikov-Teschner (FZZT) antibranes. We treat the anti-FZZT branes as dynamical objects, taking their back-reaction into account. We construct the entanglement entropy from the dual matrix model and study its behavior as a function of the ’t Hooft coupling tt proportional to the number of branes, which plays the role of time. By numerical computation we observe that the entropy first increases and then decreases as tt grows, reproducing the well-known behavior of the Page curve of an evaporating black hole. The system finally exhibits a phase transition, which may be viewed as the end of the evaporation. We study the critical behavior of the entropy near the phase transition. We also make a conjecture about the late-time monotonically decreasing behavior of the entropy. We prove it in a certain limit as well as give an intuitive explanation by means of the dual matrix model.

1 Introduction

The black hole information paradox has been a long-standing puzzle in the study of quantum gravity Hawking:1976ra . In particular, the growing behavior of the entropy of thermal radiation based on Hawking’s calculation Hawking:1975vcx apparently contradicts with the unitarity of the quantum mechanics which requires that the black hole stays in a pure state. For an evaporating black hole, the Page curve Page:1993wv , a plot of the entanglement entropy of the Hawking radiation as a function of time, should show a decreasing behavior toward the end of evaporation. Recent studies revealed that the gravitational path integral receives, even semi-classically, contributions from saddle-points other than the classical black hole solution, namely the replica wormholes Penington:2019npb ; Almheiri:2019psf . This is a key to understand how the Page curve is obtained in an expected form, which partly resolves the information paradox. The idea was refined in the form of the island formula Almheiri:2019hni , which was first derived by means of holography Ryu:2006bv ; Hubeny:2007xt ; Lewkowycz:2013nqa ; Barrella:2013wja ; Faulkner:2013ana ; Engelhardt:2014gca ; Penington:2019npb ; Almheiri:2019psf ; Almheiri:2019hni and then consolidated by directly evaluating the gravitational path integral in quantum gravity in two dimensions Penington:2019kki ; Almheiri:2019qdq . See Almheiri:2020cfm for a recent review and references therein.

In Penington:2019kki the Page curve was studied by using Jackiw-Teitelboim (JT) gravity Jackiw:1984je ; Teitelboim:1983ux with the end-of-the-world (EOW) branes.111A classification of branes in JT gravity is found in Goel:2020yxl . Roughly speaking, the system is viewed as a generalization of the original Page’s model Page:1993df (see Okuyama:2021ylc for recent exact results). Page’s calculation starts with a random pure state in the bipartite Hilbert space consisting of two subspaces that represent the interior and exterior of a black hole. Taking ensemble average of the state in either of the subspaces, one obtains the reduced density matrix, from which the entanglement entropy is calculated. In Page’s model the ensemble is Gaussian in both subspaces. In the case of JT gravity with EOW branes, the ensemble in the interior is Gaussian whereas the average in the exterior is described by the double-scaled matrix integral of JT gravity Saad:2019lba . The size of the interior subspace, which is identified with the number of branes, plays the role of time.

In this paper we propose another simple toy model to understand the Page curve: JT gravity with Fateev-Zamolodchikov-Zamolodchikov-Teschner (FZZT) antibranes Fateev:2000ik ; Teschner:2000md . Our model is a simplified variant of the model of Penington:2019kki , with the EOW branes replaced by anti-FZZT branes. In our previous paper Okuyama:2021eju we showed that the matrix model description of the EOW brane in Gao:2021uro corresponds to that of a collection of infinitely many anti-FZZT branes with a particular set of parameters. It is therefore simpler to consider JT gravity with a single kind of anti-FZZT branes.

Despite this simplification, our model captures several features of black hole entropy. Most notably, the entanglement entropy exhibits the late-time decreasing behavior which is characteristic of an evaporating black hole.222The Page curve of an evaporating black hole in JT gravity was studied in different approaches Goto:2020wnk ; Cadoni:2021ypx . To reproduce this decreasing behavior, it is crucial to treat branes as dynamical objects. In the previous studies, branes are treated as either dynamical Gao:2021uro or non-dynamical Penington:2019kki . We will see from numerical computation that the late-time decreasing behavior of the entropy is reproduced only when we treat anti-FZZT branes as dynamical objects. In fact, we consider the ’t Hooft limit, in which the back-reaction of branes is not negligible and one has to treat branes as dynamical objects. We will also study how this decreasing behavior arises from the viewpoint of the matrix model and make a conjecture about monotonicity, which we will prove in a certain limit.

Our model exhibits a phase transition as “time” grows. The transition may be viewed as the end of the evaporation of black hole. We will study the critical behavior of the entropy near the transition point. One can consider JT gravity with FZZT branes and study the Page curve in the same manner. In this case, however, no phase transition occurs and the entropy continue increasing. All these results derived from the matrix model are in perfect accordance with the semi-classical analysis on the gravity side: Dilaton gravities with nontrivial dilaton potential were studied as deformations of JT gravity Maxfield:2020ale ; Witten:2020wvy and black hole solutions in these gravities were also studied Witten:2020ert . JT gravity with (anti-)FZZT branes can be viewed as this type of dilaton gravity Okuyama:2021eju . We will study its black hole solutions and see the continuous growth of the entropy in the FZZT setup and the phase transition in the anti-FZZT setup. Thus, in this paper we concentrate on the case of anti-FZZT branes.

This paper is organized as follows. In section 2, we describe our model and explain the general method of computing the entropy of the Hawking radiation. In section 3, we explain how the phase transition occurs and study the critical behavior of the entropy. In section 4, we numerically study the Page curve, i.e. the evolution of the entropy as a function of the ’t Hooft coupling. We also make a conjecture about the late-time monotonically decreasing behavior of the entropy. In section 5, we prove the conjecture in a certain limit. We also give an intuitive explanation of the reason why the entropy decreases. In section 6, we study black hole solutions from the viewpoint of dilaton gravity. Finally we conclude in section 7. In appendix A, we give a derivation of the Schwinger-Dyson equation (22) based on the saddle point method.

2 Entropy of radiation from dynamical anti-FZZT branes

In this section we will describe our model and explain the general method of computing the entropy of the Hawking radiation. In many parts of our formulation we follow the method of Penington:2019kki with EOW branes being replaced by anti-FZZT branes. In our study, however, branes are treated as dynamical objects. This is along the lines of Gao:2021uro and an important difference from Penington:2019kki .

2.1 Matrix integral and black hole microstates

Let us consider general 2d topological gravity with KK dynamical anti-FZZT branes.333In this paper we will eventually restrict ourselves to the JT gravity case, but most parts of our formalism can be applied to other 2d gravities as well. It is described by the double scaling limit of the matrix integral

Z\displaystyle Z =𝑑HeTrV(H)det(ξ+H)K\displaystyle=\int dHe^{-\operatorname{Tr}V(H)}\det(\xi+H)^{-K} (1)
=𝑑H𝑑Q𝑑QeTrV(H)TrQ(ξ+H)Q.\displaystyle=\int dHdQdQ^{\dagger}e^{-\operatorname{Tr}V(H)-\operatorname{Tr}Q^{\dagger}(\xi+H)Q}.

Here HH and QQ are N×NN\times N hermitian and N×KN\times K complex matrices respectively. ξ\xi is a parameter characterizing the anti-FZZT brane, which is now taken to be common to all KK branes. The potential could have been normalized as

1gsV(H),\displaystyle\frac{1}{g_{\rm s}}V(H), (2)

where gsg_{\rm s} is the genus counting parameter, so that the genus expansion is manifest. In this paper we include gs1g_{\rm s}^{-1} in VV for simplicity. In the double scaling limit, NN is sent to infinity and the potential turns into the effective potential. In this paper we will further take the ’t Hooft limit

K,gs0withtgsKfixed\displaystyle K\to\infty,~{}g_{\rm s}\to 0\quad\text{with}~{}~{}t\equiv g_{\rm s}K~{}~{}\text{fixed} (3)

and evaluate quantities in the planar approximation. That is, we will ignore all higher-order corrections of expansions in gsg_{\rm s} and K1K^{-1}.

The matrices H,QH,Q are often denoted by their components HabH_{ab}, QaiQ_{ai}, where a,b=1,,Na,b=1,\ldots,N are “color” indices and i,j=1,,Ki,j=1,\ldots,K are “flavor” indices. The color degrees of freedom are used for describing bulk gravity while the flavor degrees of freedom are thought of as describing the interior partners of the early Hawking radiation. One can regard the matrix element HabH_{ab} as

Hab=a|H|b,\displaystyle H_{ab}=\langle a|H|b\rangle, (4)

where HH is a Hamiltonian operator and {|a}a=1N\{|a\rangle\}_{a=1}^{N} form an orthonormal basis of the corresponding NN dimensional Hilbert space

a|b=δab,1=a|aa|.\displaystyle\langle a|b\rangle=\delta_{ab},\qquad 1=\sum_{a}|a\rangle\langle a|. (5)

For iith random vector variable QaiQ_{ai} we consider the (canonical) thermal pure quantum state Sugiura:2013pla ; Goto:2021mbt

|ψi=ae12βH|aQai=a,b|b(e12βH)baQai.\displaystyle|\psi_{i}\rangle=\sum_{a}e^{-\frac{1}{2}\beta H}|a\rangle Q_{ai}=\sum_{a,b}|b\rangle(e^{-\frac{1}{2}\beta H})_{ba}Q_{ai}. (6)

Here β\beta is the inverse temperature, which is identified with the (renormalized) length of an asymptotic boundary in 2d gravity. |ψi|\psi_{i}\rangle play the role of the black hole microstates.

2.2 Ensemble average

To study the entropy, we will compute the average of overlaps such as ψi|ψj\langle\psi_{i}|\psi_{j}\rangle. We define the average of 𝒪{\cal O} by

𝒪¯=𝑑H𝑑Q𝑑QeTrV(H)TrQ(ξ+H)Q𝒪.\langle\overline{{\cal O}}\rangle=\int dHdQdQ^{\dagger}e^{-\operatorname{Tr}V(H)-\operatorname{Tr}Q^{\dagger}(\xi+H)Q}{\cal O}. (7)

Here the angle brackets \langle\ \rangle represent averaging over the color degrees of freedom while the overline  ¯\overline{\rule{10.00002pt}{0.0pt}\rule{0.0pt}{6.45831pt}} represents averaging over the flavor degrees of freedom. It is convenient to change the variable as

Q=(ξ+H)12C,\displaystyle Q=(\xi+H)^{-\frac{1}{2}}C, (8)

so that the new random variable CC obeys the Gaussian distribution

𝒪¯=𝑑H𝑑C𝑑Cdet(ξ+H)KeTrV(H)TrCC𝒪.\langle\overline{{\cal O}}\rangle=\int dHdCdC^{\dagger}\det(\xi+H)^{-K}e^{-\operatorname{Tr}V(H)-\operatorname{Tr}C^{\dagger}C}{\cal O}. (9)

Thus, in terms of CC the flavor average becomes nothing but the Gaussian average. Note that the determinant factor is recovered from the integration measure. On the other hand, the thermal pure quantum state (6) becomes (see also appendix D of Penington:2019kki )

|ψi=a,b|b[e12βH(ξ+H)12]baCai.\displaystyle|\psi_{i}\rangle=\sum_{a,b}|b\rangle\bigl{[}e^{-\frac{1}{2}\beta H}(\xi+H)^{-\frac{1}{2}}\bigr{]}_{ba}C_{ai}. (10)

For our discussion it is convenient to express (10) as

|ψi=a,b|b(A)baCai\displaystyle|\psi_{i}\rangle=\sum_{a,b}|b\rangle(\sqrt{A})_{ba}C_{ai} (11)

with

A(H)=eβHξ+H.\displaystyle A(H)=\frac{e^{-\beta H}}{\xi+H}. (12)

We then consider the overlaps such as

Wij\displaystyle W_{ij} ψi|ψj=a,bAabCaiCbj,\displaystyle\equiv\langle\psi_{i}|\psi_{j}\rangle=\sum_{a,b}A_{ab}C_{ai}^{*}C_{bj}, (13)
WijWji\displaystyle W_{ij}W_{ji} =|ψi|ψj|2=a,b,a,bAabAbaCaiCbjCaiCbj.\displaystyle=|\langle\psi_{i}|\psi_{j}\rangle|^{2}=\sum_{a,b,a^{\prime},b^{\prime}}A_{ab}A_{b^{\prime}a^{\prime}}C_{ai}^{*}C_{bj}C_{a^{\prime}i}C_{b^{\prime}j}^{*}.

Recall that the Gaussian average of CC can be computed by the Wick contraction

CaiCbj¯\displaystyle\overline{C_{ai}^{*}C_{bj}} =δabδij,\displaystyle=\delta_{ab}\delta_{ij}, (14)
CaiCbjCaiCbj¯\displaystyle\overline{C_{ai}^{*}C_{bj}C_{a^{\prime}i}C_{b^{\prime}j}^{*}} =δijδabδab+δaaδbb.\displaystyle=\delta_{ij}\delta_{ab}\delta_{a^{\prime}b^{\prime}}+\delta_{aa^{\prime}}\delta_{bb^{\prime}}.

By using these formulas, the average of the overlaps (13) are given by

ψi|ψj¯\displaystyle\overline{\langle\psi_{i}|\psi_{j}\rangle} =δijTrA,\displaystyle=\delta_{ij}\operatorname{Tr}A, (15)
|ψi|ψj|2¯\displaystyle\overline{|\langle\psi_{i}|\psi_{j}\rangle|^{2}} =δij(TrA)2+TrA2.\displaystyle=\delta_{ij}(\operatorname{Tr}A)^{2}+\operatorname{Tr}A^{2}.

As discussed in Penington:2019kki , one can visualize the above computation (15) by drawing diagrams. For instance, ψi|ψj\langle\psi_{i}|\psi_{j}\rangle in (13) can be represented by the following diagram

ψi|ψj=(CAC)ij=ij(C)iaAabCbj.\displaystyle\leavevmode\hbox to271.14pt{\vbox to29.92pt{\pgfpicture\makeatletter\hbox{\hskip 104.9894pt\lower-11.78091pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{{}} {}{{}}{}{{{}} {}{}{}{}{}{}{}{} }{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@moveto{1.13815pt}{0.0pt}\pgfsys@curveto{1.13815pt}{0.62859pt}{0.62859pt}{1.13815pt}{0.0pt}{1.13815pt}\pgfsys@curveto{-0.62859pt}{1.13815pt}{-1.13815pt}{0.62859pt}{-1.13815pt}{0.0pt}\pgfsys@curveto{-1.13815pt}{-0.62859pt}{-0.62859pt}{-1.13815pt}{0.0pt}{-1.13815pt}\pgfsys@curveto{0.62859pt}{-1.13815pt}{1.13815pt}{-0.62859pt}{1.13815pt}{0.0pt}\pgfsys@closepath\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ } {}{{}}{} {}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setdash{3.0pt,3.0pt}{0.0pt}\pgfsys@invoke{ }{}\pgfsys@moveto{1.13815pt}{0.0pt}\pgfsys@lineto{45.52371pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{{}}{}{{}}{}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.8pt}\pgfsys@invoke{ }{}\pgfsys@moveto{45.52371pt}{0.0pt}\pgfsys@curveto{68.28557pt}{11.38092pt}{102.42836pt}{-11.38092pt}{113.8093pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{{}}{} {}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setdash{3.0pt,3.0pt}{0.0pt}\pgfsys@invoke{ }{}\pgfsys@moveto{113.8093pt}{0.0pt}\pgfsys@lineto{160.47118pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{{}}{}{{{}} {}{}{}{}{}{}{}{} }{}\pgfsys@moveto{160.47118pt}{0.0pt}\pgfsys@moveto{161.60933pt}{0.0pt}\pgfsys@curveto{161.60933pt}{0.62859pt}{161.09976pt}{1.13815pt}{160.47118pt}{1.13815pt}\pgfsys@curveto{159.84259pt}{1.13815pt}{159.33302pt}{0.62859pt}{159.33302pt}{0.0pt}\pgfsys@curveto{159.33302pt}{-0.62859pt}{159.84259pt}{-1.13815pt}{160.47118pt}{-1.13815pt}\pgfsys@curveto{161.09976pt}{-1.13815pt}{161.60933pt}{-0.62859pt}{161.60933pt}{0.0pt}\pgfsys@closepath\pgfsys@moveto{160.47118pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ } {}{{}}{}{{}}{}\pgfsys@moveto{-11.38092pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ }\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{{}{}}}{{}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-101.65639pt}{-2.9211pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$\langle\psi_{i}|\psi_{j}\rangle=(C^{\dagger}AC)_{ij}=$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {}{{}}{}{{}}{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ }\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-1.72256pt}{3.533pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$i$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {}{{}}{}{{}}{}\pgfsys@moveto{160.47118pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ }\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{158.12593pt}{5.47743pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$j$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {}{{}}{}{{}}{}\pgfsys@moveto{22.76186pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ }\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{11.25258pt}{6.033pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$(C^{\dagger})_{ia}$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {}{{}}{}{{}}{}\pgfsys@moveto{79.6665pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ }\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{73.23479pt}{5.47743pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$A_{ab}$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {}{{}}{}{{}}{}\pgfsys@moveto{136.57115pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ }\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{130.12491pt}{6.56633pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$C_{bj}$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}. (16)

The black thick curve labeled by the color matrix AabA_{ab} corresponds to the asymptotic boundary of 2d spacetime while the dashed lines correspond to the flavor degrees of freedom C,CC,C^{\dagger}. The gravitational path integral in the presence of branes is given by the matrix integral (9). One can easily see that the gravitational computations in eq. (2.10) and Figure 3 of Penington:2019kki agree with the first and the second lines of (15), respectively.

2.3 Reduced density matrix of radiation

As explained in Penington:2019kki , the reduced density matrix of radiation is represented by the ensemble average of

ϱij=Wiji=1KWii=(WTrW)ij.\displaystyle\varrho_{ij}=\frac{W_{ij}}{\sum_{i=1}^{K}W_{ii}}=\left(\frac{W}{\operatorname{Tr}W}\right)_{ij}. (17)

This is normalized as Trϱ=1\operatorname{Tr}\varrho=1. Let us first consider the “purity” Trϱ2¯\overline{\operatorname{Tr}\varrho^{2}} as an example. In the planar approximation, we can take the average of the numerator and the denominator of Trϱ2\operatorname{Tr}\varrho^{2} independently

Trϱ2¯TrW2¯TrW¯2\displaystyle\overline{\operatorname{Tr}\varrho^{2}}\approx\frac{\overline{\operatorname{Tr}W^{2}}}{\overline{\operatorname{Tr}W}^{2}} =K(TrA)2+K2TrA2(KTrA)2\displaystyle=\frac{K(\operatorname{Tr}A)^{2}+K^{2}\operatorname{Tr}A^{2}}{(K\operatorname{Tr}A)^{2}} (18)
=1K+TrA2(TrA)2.\displaystyle=\frac{1}{K}+\frac{\operatorname{Tr}A^{2}}{(\operatorname{Tr}A)^{2}}.

Similarly, the average of Trϱn\operatorname{Tr}\varrho^{n} is approximated as

Trϱn¯TrWn¯(TrW¯)n=TrWn¯(KTrA)n.\displaystyle\overline{\operatorname{Tr}\varrho^{n}}\approx\frac{\overline{\operatorname{Tr}W^{n}}}{(\overline{\operatorname{Tr}W})^{n}}=\frac{\overline{\operatorname{Tr}W^{n}}}{(K\operatorname{Tr}A)^{n}}. (19)

TrWn¯\overline{\operatorname{Tr}W^{n}} in the numerator can be computed by using the Wick contraction of CC and CC^{\dagger}. In the planar approximation one obtains

TrW¯\displaystyle\overline{\operatorname{Tr}W} =KTrA,\displaystyle=K\operatorname{Tr}A, (20)
TrW2¯\displaystyle\overline{\operatorname{Tr}W^{2}} =K(TrA)2+K2TrA2,\displaystyle=K(\operatorname{Tr}A)^{2}+K^{2}\operatorname{Tr}A^{2},
TrW3¯\displaystyle\overline{\operatorname{Tr}W^{3}} =K(TrA)3+3K2TrA2TrA+K3TrA3,\displaystyle=K(\operatorname{Tr}A)^{3}+3K^{2}\operatorname{Tr}A^{2}\operatorname{Tr}A+K^{3}\operatorname{Tr}A^{3},
TrW4¯\displaystyle\overline{\operatorname{Tr}W^{4}} =K(TrA)4+6K2TrA2(TrA)2+2K3(TrA2)2+4K3TrA3TrA+K4TrA4.\displaystyle=K(\operatorname{Tr}A)^{4}+6K^{2}\operatorname{Tr}A^{2}(\operatorname{Tr}A)^{2}+2K^{3}(\operatorname{Tr}A^{2})^{2}+4K^{3}\operatorname{Tr}A^{3}\operatorname{Tr}A+K^{4}\operatorname{Tr}A^{4}.

In fact, TrWn¯\overline{\operatorname{Tr}W^{n}} can be computed efficiently by means of the generating function

R(λ)=Tr1λϱ=n=0Trϱnλn+1=Kλ+n=1TrWn¯λn+1(KTrA)n.\displaystyle R(\lambda)=\operatorname{Tr}\frac{1}{\lambda-\varrho}=\sum_{n=0}^{\infty}\frac{\operatorname{Tr}\varrho^{n}}{\lambda^{n+1}}=\frac{K}{\lambda}+\sum_{n=1}^{\infty}\frac{\overline{\operatorname{Tr}W^{n}}}{\lambda^{n+1}(K\operatorname{Tr}A)^{n}}. (21)

In the planar approximation R(λ)R(\lambda) satisfies

λR(λ)=K+n=1R(λ)nTrAn(KTrA)n.\displaystyle\lambda R(\lambda)=K+\sum_{n=1}^{\infty}\frac{R(\lambda)^{n}\operatorname{Tr}A^{n}}{(K\operatorname{Tr}A)^{n}}. (22)

This equation was derived diagrammatically in Penington:2019kki . We give an alternative derivation based on the saddle point method in appendix A. By plugging (21) into (22), TrWn¯\overline{\operatorname{Tr}W^{n}} can be obtained recursively. In this way, in the planar approximation we obtain

Trϱn¯1Kn1+n(n1)2Kn2TrA2TrA2++nKTrAn1TrAn1+TrAnTrAn.\displaystyle\langle\overline{\operatorname{Tr}\varrho^{n}}\rangle\approx\frac{1}{K^{n-1}}+\frac{n(n-1)}{2K^{n-2}}\frac{\langle\operatorname{Tr}A^{2}\rangle}{\langle\operatorname{Tr}A\rangle^{2}}+\cdots+\frac{n}{K}\frac{\langle\operatorname{Tr}A^{n-1}\rangle}{\langle\operatorname{Tr}A\rangle^{n-1}}+\frac{\langle\operatorname{Tr}A^{n}\rangle}{\langle\operatorname{Tr}A\rangle^{n}}. (23)

Thus, to compute Trϱn¯\langle\overline{\operatorname{Tr}\varrho^{n}}\rangle we need to evaluate

TrAn=𝑑HeTrV(H)det(ξ+H)KTrAn.\displaystyle\langle\operatorname{Tr}A^{n}\rangle=\int dHe^{-\operatorname{Tr}V(H)}\det(\xi+H)^{-K}\operatorname{Tr}A^{n}. (24)

We evaluate it in the double scaling limit. In the planar approximation we have only to consider the genus zero part. It can be expressed in terms of the leading-order density ρ0(E)\rho_{0}(E) of the eigenvalues of HH. As we studied in Okuyama:2021eju , for Witten-Kontsevich topological gravity with general couplings {tk}(k0)\{t_{k}\}\ (k\in{\mathbb{Z}}_{\geq 0}), ρ0(E)\rho_{0}(E) is given by

ρ0(E)=12πgsE0EdvEvf(v)(v)\displaystyle\rho_{0}(E)=\frac{1}{\sqrt{2}\pi g_{\rm s}}\int_{E_{0}}^{E}\frac{dv}{\sqrt{E-v}}\frac{\partial f(-v)}{\partial(-v)} (25)

with

f(u):=k=0(δk,1tk)ukk!.\displaystyle f(u):=\sum_{k=0}^{\infty}(\delta_{k,1}-t_{k})\frac{u^{k}}{k!}. (26)

The threshold energy E0E_{0} is determined by the condition (the genus-zero string equation)

f(E0)=0.\displaystyle f(-E_{0})=0. (27)

In this paper we consider JT gravity, which corresponds to a particular background tk=γkt_{k}=\gamma_{k} with Mulase:2006baa ; Dijkgraaf:2018vnm ; Okuyama:2019xbv 444Another way to obtain JT gravity is to take the pp\to\infty limit of the (2,p)(2,p) minimal string Seiberg:2019 ; Saad:2019lba . Entanglement entropy in this context was studied recently in Hirano:2021rzg .

γ0=γ1=0,γk=(1)k(k1)!(k2).\displaystyle\gamma_{0}=\gamma_{1}=0,\quad\gamma_{k}=\frac{(-1)^{k}}{(k-1)!}\quad(k\geq 2). (28)

As we studied in Okuyama:2021eju , the effect of anti-FZZT branes, i.e. the insertion of det(ξ+H)K\det(\xi+H)^{-K} amounts to shifting the couplings tkt_{k} of topological gravity as555This shift of couplings was first recognized in the theory of soliton equations Date:1982yeu and appears in various contexts of matrix models and related subjects. For more details, see Okuyama:2021eju and references therein.

tk=γk+t(2k1)!!(2ξ)k12.\displaystyle t_{k}=\gamma_{k}+t(2k-1)!!(2\xi)^{-k-\frac{1}{2}}. (29)

This is valid as long as Reξ>0{\rm Re}\,\xi>0. Here tt is the ’t Hooft coupling in (3). Thus, (24) is evaluated as

TrAnTrAng=0=E0𝑑Eρ0(E)A(E)n,Zn,\displaystyle\begin{aligned} \langle\operatorname{Tr}A^{n}\rangle&\approx\langle\operatorname{Tr}A^{n}\rangle^{g=0}\\ &=\int_{E_{0}}^{\infty}dE\rho_{0}(E)A(E)^{n},\\ &\equiv Z_{n},\end{aligned} (30)

where ρ0(E)\rho_{0}(E) is now evaluated in the background (29). In this background, (26) becomes

f(u=v)f(v,t)=vI1(2v)+t2ξ+2v,\displaystyle f(u=-v)\equiv f(v,t)=\sqrt{v}I_{1}(2\sqrt{v})+\frac{t}{\sqrt{2\xi+2v}}, (31)

where we have changed the variable as v=uv=-u for convenience and Ik(z)I_{k}(z) denotes the modified Bessel function of the first kind. (25) then becomes

ρ0(E)=12πgs(E0E𝑑vI0(2v)EvtE+ξEE02(E0+ξ)).\displaystyle\rho_{0}(E)=\frac{1}{\sqrt{2}\pi g_{\rm s}}\left(\int_{E_{0}}^{E}dv\frac{I_{0}(2\sqrt{v})}{\sqrt{E-v}}-\frac{t}{E+\xi}\sqrt{\frac{E-E_{0}}{2(E_{0}+\xi)}}\right). (32)

Note that in Okuyama:2021eju we calculated ρ0(E)\rho_{0}(E) for JT gravity in the presence of KK FZZT branes. The above ρ0(E)\rho_{0}(E) for anti-FZZT branes is essentially identical to this except the sign of the ’t Hooft coupling tt. Note also that this expression of ρ0(E)\rho_{0}(E) is valid as long as tt is not greater than the critical value tct_{\rm c}. We will explain this in section 3.1.

We emphasize that we have treated anti-FZZT branes as dynamical objects. More specifically, in (24) the color average is evaluated in the presence of the determinant factor det(ξ+H)K\det(\xi+H)^{-K} and as a consequence the deformed eigenvalue density (32) is used in (30). This is the main difference from the approach of Penington:2019kki , which is based on the probe brane approximation at genus-zero

TrAnprobe|g=0\displaystyle\langle\operatorname{Tr}A^{n}\rangle_{\rm probe}\Big{|}_{g=0} =𝑑HeTrV(H)TrAn|g=0\displaystyle=\int dHe^{-\operatorname{Tr}V(H)}\operatorname{Tr}A^{n}\Big{|}_{g=0} (33)
=0𝑑Eρ0JT(E)A(E)n,\displaystyle=\int_{0}^{\infty}dE\rho_{0}^{\text{JT}}(E)A(E)^{n},

where the original JT gravity density of state ρ0JT(E)\rho_{0}^{\text{JT}}(E) is given by

ρ0JT(E)=sinh(2E)2πgs.\displaystyle\rho_{0}^{\text{JT}}(E)=\frac{\sinh(2\sqrt{E})}{\sqrt{2}\pi g_{\rm s}}. (34)

However, in Penington:2019kki the same ’t Hooft limit as ours (3) is used. As we argued in Okuyama:2021eju , in this limit the back-reaction of (anti-)FZZT branes cannot be ignored and the couplings tkt_{k} are shifted due to the insertion of branes. As a consequence, the eigenvalue density is deformed from ρ0JT(E)\rho_{0}^{\text{JT}}(E) in (34) to ρ0(E)\rho_{0}(E) in (32). Thus we have to use the dynamical brane picture in this limit.

2.4 Resolvent of reduced density matrix and entropy

We saw in the last subsection that the ensemble averages of ϱ\varrho in the planar approximation are expressed in terms of ZnZ_{n} in (30). On the other hand, the general expression (23) of Trϱn¯\langle\overline{\operatorname{Tr}\varrho^{n}}\rangle is rather complicated as a function of nn and it is difficult to apply the replica trick directly to (23) to calculate the entropy. Instead, as detailed in Penington:2019kki , we can study the entropy using the resolvent R(λ)R(\lambda) for ϱ\varrho in (21). By substituting (30), the Schwinger-Dyson equation (22) for R(λ)R(\lambda) becomes

λR(λ)\displaystyle\lambda R(\lambda) =K+n=1R(λ)n(KZ1)nE0𝑑Eρ0(E)A(E)n\displaystyle=K+\sum_{n=1}^{\infty}\frac{R(\lambda)^{n}}{(KZ_{1})^{n}}\int_{E_{0}}^{\infty}dE\rho_{0}(E)A(E)^{n} (35)
=K+E0𝑑Eρ0(E)w(E)R(λ)Kw(E)R(λ),\displaystyle=K+\int_{E_{0}}^{\infty}dE\rho_{0}(E)\frac{w(E)R(\lambda)}{K-w(E)R(\lambda)},

where we have defined

w(E)=A(E)Z1.\displaystyle w(E)=\frac{A(E)}{Z_{1}}. (36)

Following Penington:2019kki , we divide the integral in (35) into two pieces

λR(λ)\displaystyle\lambda R(\lambda) =K+E0EK𝑑Eρ0(E)w(E)R(λ)Kw(E)R(λ)+EK𝑑Eρ0(E)w(E)R(λ)Kw(E)R(λ)\displaystyle=K+\int_{E_{0}}^{E_{K}}dE\rho_{0}(E)\frac{w(E)R(\lambda)}{K-w(E)R(\lambda)}+\int_{E_{K}}^{\infty}dE\rho_{0}(E)\frac{w(E)R(\lambda)}{K-w(E)R(\lambda)} (37)
K+E0EK𝑑Eρ0(E)w(E)R(λ)Kw(E)R(λ)+λ0R(λ),\displaystyle\approx K+\int_{E_{0}}^{E_{K}}dE\rho_{0}(E)\frac{w(E)R(\lambda)}{K-w(E)R(\lambda)}+\lambda_{0}R(\lambda),

where λ0\lambda_{0} and EKE_{K} are defined by

λ0\displaystyle\lambda_{0} =1KEK𝑑Eρ0(E)w(E),\displaystyle=\frac{1}{K}\int_{E_{K}}^{\infty}dE\rho_{0}(E)w(E), (38)
K\displaystyle K =E0EK𝑑Eρ0(E).\displaystyle=\int_{E_{0}}^{E_{K}}dE\rho_{0}(E).

By rewriting (37) as

R(λ)=Kλλ0+1λλ0E0EK𝑑Eρ0(E)w(E)R(λ)Kw(E)R(λ),\displaystyle R(\lambda)=\frac{K}{\lambda-\lambda_{0}}+\frac{1}{\lambda-\lambda_{0}}\int_{E_{0}}^{E_{K}}dE\rho_{0}(E)\frac{w(E)R(\lambda)}{K-w(E)R(\lambda)}, (39)

we can solve R(λ)R(\lambda) by the iteration starting from R(λ)=K/(λλ0)R(\lambda)=K/(\lambda-\lambda_{0}). As discussed in Penington:2019kki , the second order iteration gives

R(λ)\displaystyle R(\lambda) Kλλ0+1λλ0E0EK𝑑Eρ0(E)w(E)K(λλ0)1Kw(E)K(λλ0)1\displaystyle\approx\frac{K}{\lambda-\lambda_{0}}+\frac{1}{\lambda-\lambda_{0}}\int_{E_{0}}^{E_{K}}dE\rho_{0}(E)\frac{w(E)K(\lambda-\lambda_{0})^{-1}}{K-w(E)K(\lambda-\lambda_{0})^{-1}} (40)
=Kλλ0+1λλ0E0EK𝑑Eρ0(E)w(E)λλ0w(E).\displaystyle=\frac{K}{\lambda-\lambda_{0}}+\frac{1}{\lambda-\lambda_{0}}\int_{E_{0}}^{E_{K}}dE\rho_{0}(E)\frac{w(E)}{\lambda-\lambda_{0}-w(E)}.

Using (38), we find

R(λ)\displaystyle R(\lambda) =1λλ0[E0EK𝑑Eρ0(E)+E0EK𝑑Eρ0(E)w(E)λλ0w(E)]\displaystyle=\frac{1}{\lambda-\lambda_{0}}\left[\int_{E_{0}}^{E_{K}}dE\rho_{0}(E)+\int_{E_{0}}^{E_{K}}dE\rho_{0}(E)\frac{w(E)}{\lambda-\lambda_{0}-w(E)}\right] (41)
=E0EK𝑑Eρ0(E)1λλ0w(E).\displaystyle=\int_{E_{0}}^{E_{K}}dE\rho_{0}(E)\frac{1}{\lambda-\lambda_{0}-w(E)}.

The eigenvalue density D(λ)D(\lambda) of the density matrix ϱij\varrho_{ij} is obtained from the discontinuity of R(λ)R(\lambda)

D(λ)=R(λi0)R(λ+i0)2πi=E0EK𝑑Eρ0(E)δ(λλ0w(E)).\displaystyle D(\lambda)=\frac{R(\lambda-\mathrm{i}0)-R(\lambda+\mathrm{i}0)}{2\pi\mathrm{i}}=\int_{E_{0}}^{E_{K}}dE\rho_{0}(E)\delta\bigl{(}\lambda-\lambda_{0}-w(E)\bigr{)}. (42)

Finally, the von Neumann entropy is given by

S\displaystyle S =𝑑λD(λ)λlogλ\displaystyle=-\int d\lambda D(\lambda)\lambda\log\lambda (43)
=E0EK𝑑Eρ0(E)(λ0+w(E))log(λ0+w(E)).\displaystyle=-\int_{E_{0}}^{E_{K}}dE\rho_{0}(E)(\lambda_{0}+w(E))\log\bigl{(}\lambda_{0}+w(E)\bigr{)}.

We will use this expression to study the Page curve numerically in section 4.

In the rest of this section let us make several comments on the above approximation. We can check that D(λ)D(\lambda) in (42) is normalized correctly

Trϱ0=𝑑λD(λ)1=E0EK𝑑Eρ0(E)=K,\displaystyle\operatorname{Tr}\varrho^{0}=\int d\lambda D(\lambda)\cdot 1=\int_{E_{0}}^{E_{K}}dE\rho_{0}(E)=K, (44)

where we have used (38). We also find

Trϱ\displaystyle\operatorname{Tr}\varrho =𝑑λD(λ)λ=E0EK𝑑Eρ0(E)(λ0+w(E))\displaystyle=\int d\lambda D(\lambda)\cdot\lambda=\int_{E_{0}}^{E_{K}}dE\rho_{0}(E)(\lambda_{0}+w(E)) (45)
=Kλ0+E0EK𝑑Eρ0(E)w(E)\displaystyle=K\lambda_{0}+\int_{E_{0}}^{E_{K}}dE\rho_{0}(E)w(E)
=EK𝑑Eρ0(E)w(E)+E0EK𝑑Eρ0(E)w(E)\displaystyle=\int_{E_{K}}^{\infty}dE\rho_{0}(E)w(E)+\int_{E_{0}}^{E_{K}}dE\rho_{0}(E)w(E)
=E0𝑑Eρ0(E)w(E)\displaystyle=\int_{E_{0}}^{\infty}dE\rho_{0}(E)w(E)
=1Z1E0𝑑Eρ0(E)A(E)=1Z1Z1=1,\displaystyle=\frac{1}{Z_{1}}\int_{E_{0}}^{\infty}dE\rho_{0}(E)A(E)=\frac{1}{Z_{1}}Z_{1}=1,

where we have used (38).

λ0\lambda_{0} in (38) can be written as

λ0\displaystyle\lambda_{0} =1KE0𝑑Eρ0(E)w(E)1KE0EK𝑑Eρ0(E)w(E)\displaystyle=\frac{1}{K}\int_{E_{0}}^{\infty}dE\rho_{0}(E)w(E)-\frac{1}{K}\int_{E_{0}}^{E_{K}}dE\rho_{0}(E)w(E) (46)
=1Kw¯K,\displaystyle=\frac{1}{K}-\overline{w}_{K},

where we have used (45) and defined

w¯K1KE0EK𝑑Eρ0(E)w(E)=E0EK𝑑Eρ0(E)w(E)E0EK𝑑Eρ0(E).\displaystyle\overline{w}_{K}\equiv\frac{1}{K}\int_{E_{0}}^{E_{K}}dE\rho_{0}(E)w(E)=\frac{\int_{E_{0}}^{E_{K}}dE\rho_{0}(E)w(E)}{\int_{E_{0}}^{E_{K}}dE\rho_{0}(E)}. (47)

That is, w¯K\overline{w}_{K} is the average of w(E)w(E) in the “post Page” subspace E<EKE<E_{K}. Thus the resolvent is written as

R(λ)=Tr1λϱ=E0EK𝑑Eρ0(E)1λλ(E),\displaystyle R(\lambda)=\operatorname{Tr}\frac{1}{\lambda-\varrho}=\int_{E_{0}}^{E_{K}}dE\rho_{0}(E)\frac{1}{\lambda-\lambda(E)}, (48)

where

λ(E)=1K+w(E)w¯K.\displaystyle\lambda(E)=\frac{1}{K}+w(E)-\overline{w}_{K}. (49)

λ(E)\lambda(E) behaves as

λ(E){1K(Kgs1),w(E)w¯K(Kgs1).\lambda(E)\approx\left\{\begin{aligned} &\frac{1}{K}\quad&(K\ll g_{\rm s}^{-1}),\\ &w(E)-\overline{w}_{K}\quad&(K\gg g_{\rm s}^{-1}).\end{aligned}\right. (50)

This corresponds to Figure 6 in Penington:2019kki . Note that the density matrix ϱij\varrho_{ij} is originally a matrix in the flavor space, but after taking the average the spectrum λ(E)\lambda(E) of ϱij\varrho_{ij} is effectively written in terms of energy eigenvalues in the “color” space. We have to project quantities onto the “post Page” subspace E<EKE<E_{K} to ensure that the number of total state is K=E<EK𝑑Eρ0(E)K=\int_{E<E_{K}}dE\rho_{0}(E).

3 Phase transition

An interesting feature of the anti-FZZT brane background in JT gravity is that the system exhibits a phase transition as the ’t Hooft coupling tt varies. In this section we discuss this phase transition and study the critical behavior of the entropy.

3.1 Threshold energy

Let us first clarify the definition of the threshold energy E0E_{0}. In the last section we saw that E0E_{0} is determined by the threshold energy condition (27). For JT gravity with anti-FZZT branes, ff is given by (31) and the condition (27) is written explicitly as

E0I1(2E0)=t2(E0+ξ).\displaystyle\sqrt{E_{0}}I_{1}(2\sqrt{E_{0}})=-\frac{t}{\sqrt{2(E_{0}+\xi)}}. (51)

The threshold energy E0E_{0} is determined as a real solution of this equation. Here, t>0t>0 by definition and we take ξ>0\xi>0 in order for the shift of the couplings (29) to be valid. We show the plots of both sides of the equation (51) in Figure 1.

Refer to caption
Figure 1: The threshold energy E0E_{0} is determined by the equation (51). The red curve represents the graph of EI1(2E)\sqrt{E}I_{1}(2\sqrt{E}) while the blue and green curves are the graphs of t/2(E+ξ)-t/\sqrt{2(E+\xi)} with t=3t=3 and t=tc4.46t=t_{\rm c}\approx 4.46 respectively, where we set ξ=27\xi=27. As we see in this example, (51) could have multiple real solutions. The threshold energy E0E_{0} is determined as the largest real solution, as indicated by a dot. The horizontal location of the green dot gives the critical value E0cE_{0}^{\rm c}.

We see that EI1(2E)\sqrt{E}I_{1}(2\sqrt{E}) oscillates for E<0E<0 and increases monotonically for E>0E>0 starting from the origin, while t/2(E+ξ)-t/\sqrt{2(E+\xi)} is always negative. Therefore E0E_{0} has to be negative if it exists. However, the number of real solutions of (51) varies depending on the values of tt and ξ\xi. In particular, (51) has no real solution when tt is very large, whereas it has multiple real solutions when ξ\xi is large and tt is small.666See also Johnson:2020lns for a similar problem in JT gravity with conical defects. On the other hand, one can easily see that (51) always has at least one real solution for sufficiently small tt. We define E0E_{0} as the largest real solution of (51) (i.e. the solution with the smallest absolute value), so that it is continuously deformed from E0=0E_{0}=0 for the original JT gravity case t=0t=0.

As we see from Figure 1, |E0||E_{0}| is small for small tt. If we increase tt, |E0||E_{0}| also increases. Then there exists a critical point t=tct=t_{\rm c} beyond which E0E_{0} no longer continues as a real solution. Thus, we expect a phase transition. This transition is qualitatively very similar to the one discussed in Gao:2021uro in the case of EOW branes. If one continuously increases tt beyond the critical point, E0E_{0} and the second largest root turn into a pair of complex roots. It is therefore very likely that the saddle point of the matrix integral is described by an eigenvalue density with “Y” shaped support, similar to the one studied in Gao:2021uro . It would be interesting to study the model in this “Y” shaped phase further. In this paper we view this phase transition as the end of the black hole evaporation and focus on the physics before the phase transition.

3.2 Behavior of threshold energy near t=tct=t_{\rm c}

At the critical value t=tct=t_{\rm c}, the equation (51) has a double root E0cE_{0}^{\rm c}. Thus tct_{\rm c} is determined by the condition

f(E0c,tc)=0,vf(E0c,tc)=0\displaystyle f(E_{0}^{\rm c},t_{\rm c})=0,\quad\partial_{v}f(E_{0}^{\rm c},t_{\rm c})=0 (52)

with f(v,t)f(v,t) given in (31). Expanding the equation f(E0,t)=0f(E_{0},t)=0 around (v,t)=(E0c,tc)(v,t)=(E_{0}^{\rm c},t_{\rm c}), we find

0\displaystyle 0 =f(E0c,tc)+vf(E0c,tc)(E0E0c)+12v2f(E0c,tc)(E0E0c)2\displaystyle=f(E_{0}^{\rm c},t_{\rm c})+\partial_{v}f(E_{0}^{\rm c},t_{\rm c})(E_{0}-E_{0}^{\rm c})+\frac{1}{2}\partial_{v}^{2}f(E_{0}^{\rm c},t_{\rm c})(E_{0}-E_{0}^{\rm c})^{2} (53)
+tf(E0c,tc)(ttc)+.\displaystyle\hskip 10.00002pt+\partial_{t}f(E_{0}^{\rm c},t_{\rm c})(t-t_{\rm c})+\cdots.

The first two terms vanish due to (52). Thus, near t=tct=t_{\rm c} we find

E0E0cCtct(t<tc),\displaystyle E_{0}-E_{0}^{\rm c}\approx C\sqrt{t_{\rm c}-t}\quad(t<t_{\rm c}), (54)

where CC is given by

C=2tf(E0c,tc)v2f(E0c,tc).\displaystyle C=\sqrt{\frac{2\partial_{t}f(E_{0}^{\rm c},t_{\rm c})}{\partial_{v}^{2}f(E_{0}^{\rm c},t_{\rm c})}}. (55)

3.3 Effective zero-temperature entropy and von Neumann entropy near t=tct=t_{\rm c}

In Gao:2021uro , the effective zero-temperature entropy SeffS_{\text{eff}} was introduced. It is defined by the behavior of ρ0(E)\rho_{0}(E) near E=E0E=E_{0}:

ρ0(E)eSeffEE0.\displaystyle\rho_{0}(E)\sim e^{S_{\text{eff}}}\sqrt{E-E_{0}}. (56)

From (25) we find

eSeff\displaystyle e^{S_{\text{eff}}} =2πgsvf(E0,t)\displaystyle=\frac{\sqrt{2}}{\pi g_{\rm s}}\partial_{v}f(E_{0},t) (57)
=2πgs[I0(2E0)t(2ξ+2E0)32].\displaystyle=\frac{\sqrt{2}}{\pi g_{\rm s}}\left[I_{0}(2\sqrt{E_{0}})-\frac{t}{(2\xi+2E_{0})^{\frac{3}{2}}}\right].

In Figure 2, we show the plot of SeffS_{\text{eff}}. We see that SeffS_{\text{eff}} is a monotonically decreasing function of tt.

Refer to caption
Figure 2: Plot of the effective entropy SeffS_{\text{eff}}. We set ξ=18,gs=1/100\xi=18,\,g_{\rm s}=1/100 in this figure.

Near t=tct=t_{\rm c}, using (52) and (54) we find

eSeffvf(E0,t)=vf(E0c,tc)+v2f(E0c,tc)(E0E0c)+tvf(E0c,tc)(ttc)+tct.\displaystyle\begin{aligned} e^{S_{\text{eff}}}&\sim\partial_{v}f(E_{0},t)\\ &=\partial_{v}f(E_{0}^{\rm c},t_{\rm c})+\partial_{v}^{2}f(E_{0}^{\rm c},t_{\rm c})(E_{0}-E_{0}^{\rm c})+\partial_{t}\partial_{v}f(E_{0}^{\rm c},t_{\rm c})(t-t_{\rm c})+\cdots\\ &\sim\sqrt{t_{\rm c}-t}.\end{aligned} (58)

Using the above results, we can evaluate the critical behavior of the von Neumann entropy (43). It turns out that the critical behavior of the von Neumann entropy S(t)S(t) in (43) is determined by that of the eigenvalue density near E=E0E=E_{0}

ρ0(E)eSeffEE0tctEE0.\displaystyle\rho_{0}(E)\sim e^{S_{\text{eff}}}\sqrt{E-E_{0}}\sim\sqrt{t_{\rm c}-t}\sqrt{E-E_{0}}. (59)

One can show that the contribution of the EE-integral (43) away from the edge E=E0E=E_{0} is finite at t=tct=t_{\rm c}. Subtracting this finite contribution and using (59) near E=E0E=E_{0} in (43), we find

S(t)S(tc)tct.\displaystyle S(t)-S(t_{\rm c})\sim\sqrt{t_{\rm c}-t}. (60)

In the next section we will confirm this behavior numerically.

4 Numerical study of Page curve

In section 2 we saw how to calculate the entropy. In this section we will numerically study the Page curve, i.e. the time evolution of the entropy. In Page’s original calculation logK\log K is regarded as “time” Page:1993df . Since we take the ’t Hooft limit (3), we will regard logt\log t as “time.” We will plot the entropy as a function of tt rather than logt\log t, which is convenient for seeing the critical behavior discussed in section 3.

4.1 Von Neumann entropy

We consider the von Neumann entropy (43) in JT gravity in the presence of KK anti-FZZT branes. As discussed in section 2 we compute the entropy for dynamical branes, but it is interesting to compute it in the probe brane approximation as well for the sake of comparison. In the probe brane approximation, we have E0=0E_{0}=0 and the eigenvalue density is given by ρ0JT(E)\rho_{0}^{\text{JT}}(E) in (34). We show the plot of the von Neumann entropy SS in Figure 3. We see that the entropy for the dynamical brane (solid blue curve) starts to decrease relative to the probe brane case (dashed orange curve). This is very similar to the Page curve of an evaporating black hole.

Refer to caption
Figure 3: Plot of the von Neumann entropy SS in (43) as a function of t=gsKt=g_{\rm s}K. We set ξ=18,β=4,gs=1/100\xi=18,\,\beta=4,\,g_{\rm s}=1/100 in this figure. The solid blue curve is the dynamical brane case while the dashed orange curve is the probe brane case.

As we saw in the last section, we observe that the entropy exhibits the critical behavior (60).

4.2 Rényi entropy

It is also interesting to consider the Rényi entropy. The nnth Rényi entropy SnS_{n} is defined by

Trϱn¯=e(n1)Sn.\displaystyle\langle\overline{\operatorname{Tr}\varrho^{n}}\rangle=e^{-(n-1)S_{n}}. (61)

For large KK, Trϱn¯\langle\overline{\operatorname{Tr}\varrho^{n}}\rangle is dominated by the last term of (23). Thus as tt grows, SnS_{n} approaches

S~n:=1n1logZn(Z1)n.\displaystyle\widetilde{S}_{n}:=-\frac{1}{n-1}\log\frac{Z_{n}}{(Z_{1})^{n}}. (62)

Let us first consider the second Rényi entropy

S2=log(gst+Z2Z12).\displaystyle S_{2}=-\log\left(\frac{g_{\rm s}}{t}+\frac{Z_{2}}{Z_{1}^{2}}\right). (63)

In Figure 4, we show the plot of S2S_{2} as a function of tt. We can see that S2S_{2} first increases and then decreases. Near t=tct=t_{\rm c}, S2S_{2} exhibits a critical behavior

S2(t)S2(tc)tct.\displaystyle S_{2}(t)-S_{2}(t_{\rm c})\sim\sqrt{t_{\rm c}-t}. (64)

This behavior can be derived in the same way as in the case of the von Neumann entropy.

Refer to caption
Figure 4: Plot of the second Rényi entropy S2S_{2} in (63) as a function of t=gsKt=g_{\rm s}K. We set ξ=18,β=4,gs=1/100\xi=18,\,\beta=4,\,g_{\rm s}=1/100 in this figure. The solid blue curve is S2S_{2} in (63). The dashed orange curve represents S~2\widetilde{S}_{2} in (62) while the green dashed curve is S2S_{2} without taking account of the back-reaction.

We can see that the plot of S2S_{2} has a similar behavior with the Page curve of Hawking radiation from an evaporating black hole (see e.g. Figure 7 in Almheiri:2020cfm ). We regard S~2\widetilde{S}_{2} as an analogue of the thermodynamic entropy SBHS_{\text{BH}} of an evaporating black hole. From Figure 4, we see that S~2\widetilde{S}_{2} is a monotonically decreasing function of tt (represented by the dashed orange curve).

Let us next consider the third Rényi entropy

e2S3=1K2+3Z2KZ12+Z3Z13.\displaystyle e^{-2S_{3}}=\frac{1}{K^{2}}+\frac{3Z_{2}}{KZ_{1}^{2}}+\frac{Z_{3}}{Z_{1}^{3}}. (65)

As tt grows, this approaches

e2S~3=Z3Z13.\displaystyle e^{-2\widetilde{S}_{3}}=\frac{Z_{3}}{Z_{1}^{3}}. (66)

In Figure 5 we show the plot of S3S_{3} for anti-FZZT branes.

Refer to caption
Figure 5: Plot of the third Rényi entropy S3S_{3} in (65) as a function of t=gsKt=g_{\rm s}K. We set ξ=18,β=4,gs=1/100\xi=18,\,\beta=4,\,g_{\rm s}=1/100 in this figure. The solid blue curve is S3S_{3} in (65). The dashed orange curve represents S~3\widetilde{S}_{3} in (66).

We see that this is qualitatively very similar to the S2S_{2} case. In particular, S3S_{3} is bounded from above by S~3\widetilde{S}_{3}, which monotonically decreases. From Figure 4 and 5, it is natural to regard S~n\widetilde{S}_{n} in (62) as an analogue of the thermodynamic entropy SBHS_{\text{BH}} of black hole, since SBHS_{\text{BH}} decreases monotonically during the evaporation process as well

S~nSBH.\displaystyle\widetilde{S}_{n}~{}\leftrightarrow~{}S_{\text{BH}}. (67)

Based on the above numerical results, we conjecture that S~n\widetilde{S}_{n} defined in (62) is a monotonically decreasing function of tt. More specifically, we conjecture that

tS~n<0forn>1,0<t<tc.\displaystyle\partial_{t}\widetilde{S}_{n}<0\quad\mbox{for}\quad n>1,\quad 0<t<t_{\rm c}. (68)

We will study this monotonic behavior in the next section.

5 Monotonicity of 𝑺~𝒏\widetilde{S}_{n}

In this section we study the monotonically decreasing behavior of S~n\widetilde{S}_{n}. We will prove our conjecture (68) in the large ξ\xi limit. We will also discuss how to understand intuitively this monotonically decreasing behavior.

5.1 Leading-order eigenvalue density and its derivative

In this subsection let us derive some useful formulas about the leading-order eigenvalue density ρ0(E)\rho_{0}(E) in (25) for Witten-Kontsevich gravity with general couplings {tk}\{t_{k}\}, which we will use in the next subsection.

To study general Witten-Kontsevich gravity it is convenient to introduce the Itzykson-Zuber variables Itzykson:1992ya

n(u)n(u,{tk})=m=0tn+mumm!(n0).\displaystyle\mathcal{I}_{n}(u)\equiv\mathcal{I}_{n}(u,\{t_{k}\})=\sum_{m=0}^{\infty}t_{n+m}\frac{u^{m}}{m!}\quad(n\geq 0). (69)

In terms of n\mathcal{I}_{n}, f(u)f(u) in (26) is written as

f(u)\displaystyle f(u) =u0(u)\displaystyle=u-\mathcal{I}_{0}(u) (70)

and the leading-order density ρ0(E)\rho_{0}(E) in (25) becomes

ρ0(E)\displaystyle\rho_{0}(E) =12πgsE0E𝑑v11(v)Ev.\displaystyle=\frac{1}{\sqrt{2}\pi g_{\rm s}}\int_{E_{0}}^{E}dv\frac{1-\mathcal{I}_{1}(-v)}{\sqrt{E-v}}. (71)

Let us now consider the anti-FZZT brane background (29). We are interested in how the entropy evolves as the ’t Hooft coupling tt in (3) grows. Note that tt is implicitly related to E0E_{0} by the string equation (27), from which one finds

0\displaystyle 0 =tf(E0)\displaystyle=\partial_{t}f(-E_{0}) (72)
=(tE0)(1(E0)1)t0(u)|u=E0\displaystyle=(\partial_{t}E_{0})(\mathcal{I}_{1}(-E_{0})-1)-\partial_{t}\mathcal{I}_{0}(u)\Big{|}_{u=-E_{0}}
=(tE0)(1(E0)1)12ξ+2E0.\displaystyle=(\partial_{t}E_{0})(\mathcal{I}_{1}(-E_{0})-1)-\frac{1}{\sqrt{2\xi+2E_{0}}}.

By using this relation, the tt-derivative of ρ0(E)\rho_{0}(E) is calculated as

tρ0(E)\displaystyle\partial_{t}\rho_{0}(E) =(tE0)E0ρ0(E)12πgsE0E𝑑vt1(v)Ev\displaystyle=(\partial_{t}E_{0})\partial_{E_{0}}\rho_{0}(E)-\frac{1}{\sqrt{2}\pi g_{\rm s}}\int_{E_{0}}^{E}dv\frac{\partial_{t}\mathcal{I}_{1}(-v)}{\sqrt{E-v}} (73)
=tE02πgs1(E0)1EE012πgsE0E𝑑v(Ev)12(2ξ+2v)32\displaystyle=\frac{\partial_{t}E_{0}}{\sqrt{2}\pi g_{\rm s}}\frac{\mathcal{I}_{1}(-E_{0})-1}{\sqrt{E-E_{0}}}-\frac{1}{\sqrt{2}\pi g_{\rm s}}\int_{E_{0}}^{E}dv(E-v)^{-\frac{1}{2}}(2\xi+2v)^{-\frac{3}{2}}
=12πgs(EE0)(ξ+E0)12πgs(E+ξ)EE0ξ+E0\displaystyle=\frac{1}{2\pi g_{\rm s}\sqrt{(E-E_{0})(\xi+E_{0})}}-\frac{1}{2\pi g_{\rm s}(E+\xi)}\sqrt{\frac{E-E_{0}}{\xi+E_{0}}}
=12πgs(E+ξ)ξ+E0EE0.\displaystyle=\frac{1}{2\pi g_{\rm s}(E+\xi)}\sqrt{\frac{\xi+E_{0}}{E-E_{0}}}.

Note that the background (29) is written for JT gravity, but we have never used the specific values of γk\gamma_{k} in the above derivation. Therefore, (73) is in fact valid for the anti-FZZT brane background of other gravities as well.

5.2 Proof in the large ξ\xi limit

In the last section we conjectured that S~n\widetilde{S}_{n} defined in (62) is a monotonically decreasing function of tt. In this subsection let us prove this conjecture (68) at large ξ\xi. From the expression (62), we see that (68) is equivalent to

tZnnZn>tZ1Z1.\displaystyle\frac{\partial_{t}Z_{n}}{nZ_{n}}>\frac{\partial_{t}Z_{1}}{Z_{1}}. (74)

By using the property ρ0(E0)=0\rho_{0}(E_{0})=0 and the expression (73), the tt-derivative of ZnZ_{n} in (30) is calculated as

tZn=E0𝑑Etρ0(E)A(E)n=E0𝑑E12πgsξ+E0EE0enβE(E+ξ)n+1=ξ+E02πgsenβE00𝑑E~E~12(E~+E0+ξ)n1enβE~,\displaystyle\begin{aligned} \partial_{t}Z_{n}&=\int_{E_{0}}^{\infty}dE\partial_{t}\rho_{0}(E)A(E)^{n}\\ &=\int_{E_{0}}^{\infty}dE\frac{1}{2\pi g_{\rm s}}\sqrt{\frac{\xi+E_{0}}{E-E_{0}}}\frac{e^{-n\beta E}}{(E+\xi)^{n+1}}\\ &=\frac{\sqrt{\xi+E_{0}}}{2\pi g_{\rm s}}e^{-n\beta E_{0}}\int_{0}^{\infty}d\widetilde{E}\widetilde{E}^{-\frac{1}{2}}(\widetilde{E}+E_{0}+\xi)^{-n-1}e^{-n\beta\widetilde{E}},\end{aligned} (75)

where we have set E~=EE0\widetilde{E}=E-E_{0}. For large ξ\xi, (75) is evaluated as

tZn=enβE02nπβgsξn12+𝒪(ξn32).\displaystyle\begin{aligned} \partial_{t}Z_{n}&=\frac{e^{-n\beta E_{0}}}{2\sqrt{n\pi\beta}g_{\rm s}}\xi^{-n-\frac{1}{2}}+{\cal O}\left(\xi^{-n-\frac{3}{2}}\right).\end{aligned} (76)

On the other hand, by plugging (32) into (30), ZnZ_{n} is written as

Zn=E0ρ0(E)A(E)n=12πgsE0𝑑vI0(2v)v𝑑E(Ev)12(E+ξ)nenβEt2πgs(E0+ξ)12E0𝑑E(EE0)12(E+ξ)n1enβE.\displaystyle\begin{aligned} Z_{n}&=\int_{E_{0}}^{\infty}\rho_{0}(E)A(E)^{n}\\ &=\frac{1}{\sqrt{2}\pi g_{\rm s}}\int_{E_{0}}^{\infty}dvI_{0}(2\sqrt{v})\int_{v}^{\infty}dE(E-v)^{-\frac{1}{2}}(E+\xi)^{-n}e^{-n\beta E}\\ &\hskip 10.00002pt-\frac{t}{2\pi g_{\rm s}(E_{0}+\xi)^{\frac{1}{2}}}\int_{E_{0}}^{\infty}dE(E-E_{0})^{\frac{1}{2}}(E+\xi)^{-n-1}e^{-n\beta E}.\end{aligned} (77)

In the same way as in (75)–(76), the above integrals at large ξ\xi are evaluated as

Zn=12nπβgsE0𝑑vI0(2v)(v+ξ)nenβv+𝒪(ξn1)tenβE04πgs(nβ)32ξn32+𝒪(ξn52)=12nπβgsξnE0𝑑vI0(2v)enβv+𝒪(ξn1).\displaystyle\begin{aligned} Z_{n}&=\frac{1}{\sqrt{2n\pi\beta}g_{\rm s}}\int_{E_{0}}^{\infty}dvI_{0}(2\sqrt{v})(v+\xi)^{-n}e^{-n\beta v}+{\cal O}\left(\xi^{-n-1}\right)\\ &\hskip 10.00002pt-\frac{te^{-n\beta E_{0}}}{4\sqrt{\pi}g_{\rm s}(n\beta)^{\frac{3}{2}}}\xi^{-n-\frac{3}{2}}+{\cal O}\left(\xi^{-n-\frac{5}{2}}\right)\\ &=\frac{1}{\sqrt{2n\pi\beta}g_{\rm s}\xi^{n}}\int_{E_{0}}^{\infty}dvI_{0}(2\sqrt{v})e^{-n\beta v}+{\cal O}\left(\xi^{-n-1}\right).\end{aligned} (78)

Here we see that in the leading-order of the large-ξ\xi approximation the first integral in (77) is dominant and the second integral does not contribute. Indeed, from (78) one can reproduce (76) by using the relation tE0=(2ξI0(2E0))1+𝒪(ξ3/2)\partial_{t}E_{0}=-\left(\sqrt{2\xi}I_{0}(2\sqrt{E_{0}})\right)^{-1}+{\cal O}(\xi^{-3/2}), which follows from (51) or (72). Thus we obtain

tZnnZn=enβE0n2ξE0𝑑vI0(2v)enβv+𝒪(ξ32).\displaystyle\begin{aligned} \frac{\partial_{t}Z_{n}}{nZ_{n}}&=\frac{e^{-n\beta E_{0}}}{n\sqrt{2\xi}\int_{E_{0}}^{\infty}dvI_{0}(2\sqrt{v})e^{-n\beta v}}+{\cal O}\left(\xi^{-\frac{3}{2}}\right).\end{aligned} (79)

To prove (74) at large ξ\xi, it is sufficient to show that (79) monotonically increases as nn grows:

ntZnnZn>0.\displaystyle\partial_{n}\frac{\partial_{t}Z_{n}}{nZ_{n}}>0. (80)

Since tZn/nZn>0{\partial_{t}Z_{n}}/{nZ_{n}}>0,777This follows from (79) with I0(2v)=v[vI1(2v)]>0I_{0}(2\sqrt{v})=\partial_{v}[\sqrt{v}I_{1}(2\sqrt{v})]>0 for v>E0cv>E_{0}^{\rm c}, as we can see from Figure 1. this is equivalent to showing that

nlogtZnnZn>0.\displaystyle\partial_{n}\log\frac{\partial_{t}Z_{n}}{nZ_{n}}>0. (81)

The l.h.s. of (81) is rewritten as

nlogtZnnZn=βE01n+E0𝑑vI0(2v)βvenβvE0𝑑vI0(2v)enβv=E0𝑑vI0(2v)[nβ(vE0)1]enβvnE0𝑑vI0(2v)enβv=0𝑑v~I0(2v~+E0)(nβv~1)enβv~nenβE0E0𝑑vI0(2v)enβv,\displaystyle\begin{aligned} \partial_{n}\log\frac{\partial_{t}Z_{n}}{nZ_{n}}&=-\beta E_{0}-\frac{1}{n}+\frac{\int_{E_{0}}^{\infty}dvI_{0}(2\sqrt{v})\beta ve^{-n\beta v}}{\int_{E_{0}}^{\infty}dvI_{0}(2\sqrt{v})e^{-n\beta v}}\\ &=\frac{\int_{E_{0}}^{\infty}dvI_{0}(2\sqrt{v})\left[n\beta(v-E_{0})-1\right]e^{-n\beta v}}{n\int_{E_{0}}^{\infty}dvI_{0}(2\sqrt{v})e^{-n\beta v}}\\ &=\frac{\int_{0}^{\infty}d\tilde{v}I_{0}(2\sqrt{\tilde{v}+E_{0}})\left(n\beta\tilde{v}-1\right)e^{-n\beta\tilde{v}}}{ne^{n\beta E_{0}}\int_{E_{0}}^{\infty}dvI_{0}(2\sqrt{v})e^{-n\beta v}},\end{aligned} (82)

where we have set v~=vE0\tilde{v}=v-E_{0}. The denominator of the last expression in (82) is positive (see footnote 7). By renaming v~\tilde{v} as vv, the numerator is evaluated as

0𝑑vI0(2v+E0)nβvenβv0𝑑vI0(2v+E0)enβv=I0(2v+E0)venβv|0+0𝑑v(I0(2v+E0)+I1(2v+E0)v+E0v)enβv0𝑑vI0(2v+E0)enβv=0𝑑vI1(2v+E0)v+E0venβv.\displaystyle\begin{aligned} &\int_{0}^{\infty}dvI_{0}(2\sqrt{v+E_{0}})n\beta ve^{-n\beta v}-\int_{0}^{\infty}dvI_{0}(2\sqrt{v+E_{0}})e^{-n\beta v}\\ &=-I_{0}(2\sqrt{v+E_{0}})ve^{-n\beta v}\Big{|}_{0}^{\infty}+\int_{0}^{\infty}dv\left(I_{0}(2\sqrt{v+E_{0}})+\frac{I_{1}(2\sqrt{v+E_{0}})}{\sqrt{v+E_{0}}}v\right)e^{-n\beta v}\\ &\hskip 10.00002pt-\int_{0}^{\infty}dvI_{0}(2\sqrt{v+E_{0}})e^{-n\beta v}\\ &=\int_{0}^{\infty}dv\frac{I_{1}(2\sqrt{v+E_{0}})}{\sqrt{v+E_{0}}}ve^{-n\beta v}.\end{aligned} (83)

Since the integrand is positive for any E0E_{0} satisfying E0c<E0<0E_{0}^{\rm c}<E_{0}<0,888This is easily seen from the graph of EI1(2E)=(I1(2E)/E)×E\sqrt{E}I_{1}(2\sqrt{E})=\left(I_{1}(2\sqrt{E})/\sqrt{E}\right)\times E in Figure 1. (83) is positive. Thus we have proved (81). Hence (68) has been proved at large ξ\xi.

5.3 Intuitive understanding of the decreasing behavior of S~n\widetilde{S}_{n}

Beyond the large ξ\xi approximation, it does not seem easy to find a simple analytic proof of the monotonically decreasing behavior of S~n\widetilde{S}_{n}. Alternatively, in this subsection we will explain how to understand intuitively the monotonically decreasing behavior of S~n\widetilde{S}_{n}. In contrast to the proof in the last subsection, the idea we will describe does not depend on the details of the JT gravity background and thus it can be generalized to the other gravity cases as well.

The replica index nn is sometimes identified as an analogue of the inverse temperature (see e.g. Nakaguchi:2016zqi ). Here we will pursue this analogy. To do this, let us consider the change of variable from EE to \mathcal{E} given by999(86) is rewritten as β(ξ+E)eβ(ξ+E)=βe+βξ.\displaystyle\beta(\xi+E)e^{\beta(\xi+E)}=\beta e^{\mathcal{E}+\beta\xi}. (84) This is solved by the Lambert function W(z)eW(z)=zW(z)e^{W(z)}=z as β(ξ+E)=W(βe+βξ).\displaystyle\beta(\xi+E)=W\Bigl{(}\beta e^{\mathcal{E}+\beta\xi}\Bigr{)}. (85)

e=A(E)=eβEξ+E.\displaystyle e^{-\mathcal{E}}=A(E)=\frac{e^{-\beta E}}{\xi+E}. (86)

Then we find

Zn\displaystyle Z_{n} =E0𝑑Eρ0(E)A(E)n=0𝑑D()en,\displaystyle=\int_{E_{0}}^{\infty}dE\rho_{0}(E)A(E)^{n}=\int_{\mathcal{E}_{0}}^{\infty}d\mathcal{E}D(\mathcal{E})e^{-n\mathcal{E}}, (87)

where 0=logA(E0)\mathcal{E}_{0}=-\log A(E_{0}) and

D()=(E)1ρ0(E)=ξ+Eβ(ξ+E)+1ρ0(E).\displaystyle D(\mathcal{E})=\left(\frac{\partial\mathcal{E}}{\partial E}\right)^{-1}\rho_{0}(E)=\frac{\xi+E}{\beta(\xi+E)+1}\rho_{0}(E). (88)

ZnZ_{n} in (87) takes the form of the canonical partition function with inverse temperature nn and density of states D()D(\mathcal{E}). In this picture the “thermodynamic entropy” is expressed as Dong:2016fnf ; Nakaguchi:2016zqi

Stherm=(1nn)logZn.\displaystyle S_{\text{therm}}=(1-n\partial_{n})\log Z_{n}. (89)

On the other hand, as we saw in the last subsection tS~n<0\partial_{t}\widetilde{S}_{n}<0 is equivalent to (80), which is written as

ntZnnZn\displaystyle-\partial_{n}\frac{\partial_{t}Z_{n}}{nZ_{n}} =1n2tStherm<0.\displaystyle=\frac{1}{n^{2}}\partial_{t}S_{\text{therm}}<0. (90)

Therefore, the monotonically decreasing behavior of S~n\widetilde{S}_{n} is interpreted as that of the thermodynamic entropy SthermS_{\text{therm}}.

Let us list some useful properties of SthermS_{\text{therm}}:

  1. 1.

    The threshold energy 0\mathcal{E}_{0} does not contribute to SthermS_{\text{therm}}: If we define Z~n\widetilde{Z}_{n} by

    Zn=en0Z~n,Z~n=0𝑑D(+0)en,\displaystyle Z_{n}=e^{-n\mathcal{E}_{0}}\widetilde{Z}_{n},\quad\widetilde{Z}_{n}=\int_{0}^{\infty}d\mathcal{E}D(\mathcal{E}+\mathcal{E}_{0})e^{-n\mathcal{E}}, (91)

    then we find

    Stherm=(1nn)(n0+logZ~n)=(1nn)logZ~n.\displaystyle S_{\text{therm}}=(1-n\partial_{n})(-n\mathcal{E}_{0}+\log\widetilde{Z}_{n})=(1-n\partial_{n})\log\widetilde{Z}_{n}. (92)
  2. 2.

    The overall scale of Z~n\widetilde{Z}_{n} does contribute to SthermS_{\text{therm}}: If we define

    Z~n=eS(t)Zn,\displaystyle\widetilde{Z}_{n}=e^{S(t)}Z_{n}^{\prime}, (93)

    where S(t)S(t) is nn-independent, then we find

    Stherm=(1nn)(S(t)+logZn)=S(t)+(1nn)logZn.\displaystyle S_{\text{therm}}=(1-n\partial_{n})(S(t)+\log Z_{n}^{\prime})=S(t)+(1-n\partial_{n})\log Z_{n}^{\prime}. (94)

    If we further assume that ZnZ_{n}^{\prime} is tt-independent, then the monotonically decreasing behavior of S(t)S(t) implies that of SthermS_{\text{therm}}.

  3. 3.

    SthermS_{\text{therm}} is written as

    Stherm=logZ~n+n,\displaystyle S_{\text{therm}}=\log\widetilde{Z}_{n}+n\langle\mathcal{E}\rangle, (95)

    where \langle\mathcal{E}\rangle is given by

    =nlogZ~n=1Z~n0𝑑D(+0)en.\displaystyle\langle\mathcal{E}\rangle=-\partial_{n}\log\widetilde{Z}_{n}=\frac{1}{\widetilde{Z}_{n}}\int_{0}^{\infty}d\mathcal{E}D(\mathcal{E}+\mathcal{E}_{0})e^{-n\mathcal{E}}\mathcal{E}. (96)
Refer to caption
(a) ρ0(E)\rho_{0}(E) vs EE
Refer to caption
(b) ρ0(E)\rho_{0}(E) vs EE0E-E_{0}
Figure 6: Plot of ρ0(E)\rho_{0}(E) against 6(a) EE and 6(b) EE0E-E_{0} for anti-FZZT branes. We set ξ=18\xi=18 in this plot.

Let us now focus on the case of JT gravity with anti-FZZT branes. In Figure 6 we plot ρ0(E)\rho_{0}(E) in (32) for several different values of tt. Due to property 1 of SthermS_{\text{therm}}, it is convenient to plot ρ0(E)\rho_{0}(E) against EE0E-E_{0} (see Figure 6(b)) to consider the behavior of the entropy. Then we observe that the overall scale of ρ0\rho_{0} clearly decreases as tt grows.101010As we see in (88), D()D(\mathcal{E}) is proportional to ρ0(E)\rho_{0}(E) up to a prefactor. Since this prefactor is independent of tt, the graph of D()D(\mathcal{E}) against 0\mathcal{E}-\mathcal{E}_{0} also decreases as tt grows, in a similar way as ρ0(E)\rho_{0}(E) in Figure 6(b). By crude approximation, the overall scale of ρ0\rho_{0} gives that of Z~n\widetilde{Z}_{n} and thus this implies the decreasing behavior of SthermS_{\text{therm}}, as explained in property 2. More precisely, as described in property 3, SthermS_{\text{therm}} is related to Z~n\widetilde{Z}_{n} by (95). We found numerically that each individual terms logZ~n\log\widetilde{Z}_{n} and nn\langle\mathcal{E}\rangle are not necessarily monotonically decreasing functions of tt for generic values of ξ\xi, but the sum of them is always monotonically decreasing.

To summarize, we have seen that the monotonically decreasing behavior of S~n\widetilde{S}_{n} is equivalent to that of the “thermodynamic entropy” SthermS_{\text{therm}} if we regard the replica index nn as the inverse temperature. Its decreasing behavior is intuitively understood from that of the overall scale of ρ0(E)\rho_{0}(E).

It is well known that the entropy of the Hawking radiation is bounded from above by the thermodynamic entropy SBHS_{\text{BH}} of the black hole, which is given by the area of horizon in the semi-classical approximation. For an evaporating black hole, the area of horizon decreases as time passes and this explains the decreasing behavior of the Page curve. As we saw in (67), S~n\widetilde{S}_{n} (not SthermS_{\text{therm}}) corresponds to SBHS_{\text{BH}} for n>1n>1. In general, S~n\widetilde{S}_{n} and SthermS_{\text{therm}} are different quantities. However, one can easily see that S~n\widetilde{S}_{n} becomes equal to SthermS_{\text{therm}} in the limit n1n\to 1

limn1S~n=limn1Stherm.\displaystyle\lim_{n\to 1}\widetilde{S}_{n}=\lim_{n\to 1}S_{\text{therm}}. (97)

Thus the thermodynamic entropy SBHS_{\text{BH}} of black hole literally corresponds to the “thermodynamic entropy” SthermS_{\text{therm}} in the limit n1n\to 1.

6 Dilaton gravity

Recently, dilaton gravities with nontrivial dilaton potential were studied as deformations of JT gravity Maxfield:2020ale ; Witten:2020wvy . Black hole solutions in these gravities were also discussed in Witten:2020ert . JT gravity with (anti-)FZZT branes can be viewed as this type of dilaton gravity Okuyama:2021eju . In this section we will study black hole solutions from the viewpoint of dilaton gravity.111111See also Gregori:2021tvs for recent related studies.

The action of dilaton gravity is written as Witten:2020ert

I=12d2xg(ϕR+W(ϕ)).\displaystyle I=-\frac{1}{2}\int d^{2}x\sqrt{g}(\phi R+W(\phi)). (98)

We derived that in the case of JT gravity with KK (anti-)FZZT branes the dilaton potential is given by Okuyama:2021eju

W(ϕ)={2ϕ+t2ξξ+π2ϕ2e2πϕ(anti-FZZT),2ϕt2ξξ+π2ϕ2e2πϕ(FZZT).W(\phi)=\left\{\begin{aligned} &2\phi+\frac{t\sqrt{2\xi}}{\xi+\pi^{2}\phi^{2}}e^{-2\pi\phi}&\quad&(\text{anti-FZZT}),\\ &2\phi-\frac{t\sqrt{2\xi}}{\xi+\pi^{2}\phi^{2}}e^{-2\pi\phi}&\quad&(\text{FZZT}).\end{aligned}\right. (99)

The general Euclidean black hole solution is given by

ds2=A(r)dt2+dr2A(r),ϕ(r)=r,A(r)=rhr𝑑ϕW(ϕ),\displaystyle ds^{2}=A(r)dt^{2}+\frac{dr^{2}}{A(r)},\quad\phi(r)=r,\quad A(r)=\int_{r_{h}}^{r}d\phi W(\phi), (100)

where r=rhr=r_{h} is the horizon at which A(r)A(r) vanishes. This is a one-parameter family of solutions parametrized by rh=ϕhr_{h}=\phi_{h}. The value of ϕh\phi_{h} is not fixed by the equation of motion.

The entropy of this solution is given by

S=2πϕh+S0.\displaystyle S=2\pi\phi_{h}+S_{0}. (101)

The physical condition is A(r)>0A(r)>0 for r>rhr>r_{h}.

Refer to caption
(a) W(ϕ)W(\phi) for FZZT branes
Refer to caption
(b) W(ϕ)W(\phi) for anti-FZZT branes
Figure 7: Plot of W(ϕ)W(\phi) in (99) for 7(a) FZZT branes and 7(b) anti-FZZT branes. We set ξ=18\xi=18 in this plot.

For a fixed temperature TT, ϕh\phi_{h} is determined by the condition

W(ϕh)4π=T.\displaystyle\frac{W(\phi_{h})}{4\pi}=T. (102)

In Figure 7 we show the plot of W(ϕ)W(\phi). For the FZZT branes in Figure 7(a), (102) has a unique solution ϕh\phi_{h} for a given value of TT. As tt increases, ϕh\phi_{h} also increases. Thus the entropy increases as a function of tt.

On the other hand, from Figure 7(b) for the anti-FZZT branes, one can see that there are two solutions of (102) if tt is not too large. As discussed in Witten:2020ert , the stable solution with minimal free energy corresponds to the largest root ϕh\phi_{h} of (102) (see Figure 8).

4πT4\pi TW(ϕ)W(\phi)t<tct<t_{\rm c}t=tct=t_{\rm c}t>tct>t_{\rm c}ϕh\phi_{h}
Figure 8: Stability of black hole solutions in the anti-FZZT brane setup. The largest root ϕh\phi_{h} of (102) corresponds to the stable solution. The solution no longer exists for t>tct>t_{\rm c}.

The largest root ϕh\phi_{h} of (102) decreases as tt increases. This explains the decreasing behavior of the entropy (101). Beyond some critical value t=tct=t_{\rm c}, there is no solution of (102) for a given temperature. This might be interpreted that for t>tct>t_{\rm c} there is no stable black hole solutions; at t=tct=t_{\rm c} the stable black hole disappears. This suggests that, to model the black hole evaporation process, the anti-FZZT brane setup is more suitable than the FZZT brane setup.

7 Conclusion and outlook

In this paper we studied the entanglement entropy in the matrix model of JT gravity with anti-FZZT branes, which serves as a toy model of an evaporating black hole. The entanglement entropy is defined between the color and flavor sectors, which correspond respectively to bulk gravity and to the interior partners of the early Hawking radiation. We computed the entropy in the planar approximation as well as in the ’t Hooft limit. The ’t Hooft coupling tt, which is proportional to the number of branes, plays the role of time. We computed numerically the von Neumann and Rényi entropies as functions of tt. In both cases, the entropy first increases and then decreases, which is peculiar to the Page curve of an evaporating black hole. We stress that we treated the anti-FZZT branes as dynamical objects and this was crucial to reproduce the late-time decreasing behavior of the entropy, because otherwise the entropy approaches a constant value at late time in the probe brane approximation Penington:2019kki , as we saw in Figure 3.

We saw that the system exhibits a phase transition at t=tct=t_{\rm c}. This may be viewed as the end of the evaporation of the black hole. We studied the critical behavior of the entropy and derived that it scales as in (58). As tt grows toward t=tct=t_{\rm c}, the Rényi entropy becomes dominated by S~n\widetilde{S}_{n} in (62). We conjectured that S~n\widetilde{S}_{n} monotonically decreases and proved this conjecture in the large ξ\xi limit. We also gave an intuitive explanation of this decreasing behavior. We studied black hole solutions of dilaton gravity that describes JT gravity with (anti-)FZZT branes and saw the continuous growth of the entropy in the FZZT setup as well as a phase transition in the anti-FZZT setup. This suggests that the anti-FZZT brane setup is more suitable to model an evaporating black hole.

There are many interesting open questions. We have seen that our model of dynamical branes in JT gravity serves as a good toy model for an evaporating black hole. We hope that the behavior of our model beyond the phase transition t>tct>t_{\rm c} would shed light on the deep question of the unitarity in black hole evaporation, e.g. the final state proposal in Horowitz:2003he . It would be interesting to study our matrix model beyond the phase transition t>tct>t_{\rm c} along the lines of Gao:2021uro . Our analysis of the Page curve was limited to the planar approximation. It would be interesting to compute the higher genus corrections to the Page curve. More ambitiously, it would be very interesting if we can compute the Page curve of our model non-perturbatively in gsg_{\rm s}. We leave this as an interesting future problem. We can also repeat the analysis of the Petz map in Penington:2019kki using our setup of dynamical branes. It would be interesting to study how the entanglement wedge reconstruction is modified from the result of Penington:2019kki if we take account of the back-reaction of branes.

In our calculation of the Page curve, the decreasing behavior of entropy comes from the last term of (23), which is interpreted as a contribution of replica wormholes Penington:2019kki ; Almheiri:2019qdq . The appearance of the replica wormhole is closely related to the ensemble average on the boundary side of the AdS/CFT correspondence. The role of the ensemble average in the gravitational path integral is still not well-understood and there are many conceptual issues related to this problem, such as the factorization puzzle (see e.g. Marolf:2020xie ; McNamara:2020uza ; Saad:2021rcu ; Saad:2021uzi ; Blommaert:2021fob ; Heckman:2021vzx and references therein). It is believed that the Rényi entropy is a self-averaging quantity Penington:2019kki . Nonetheless, it would be interesting to see how the Page curve of our model would look like if we pick a certain member of the ensemble and do not take an average over the random matrix (see e.g. Blommaert:2021etf for a study in this direction).

Acknowledgements.
This work was supported in part by JSPS KAKENHI Grant Nos. 19K03845, 19K03856, 21H05187 and JSPS Japan-Russia Research Cooperative Program. A preliminary result of this work was presented by one of the authors (KO) in the KMI colloquium at Nagoya University on October 13, 2021.

Appendix A Schwinger-Dyson equation from saddle point approximation

In this appendix we will derive the Schwinger-Dyson equation (22) based on the saddle point method.

Let us consider the integral

i=1Kdϕidϕieijϕi(λδijϱij)ϕj.\displaystyle\int\prod_{i=1}^{K}d\phi^{\dagger}_{i}d\phi_{i}e^{-\sum_{ij}\phi_{i}^{\dagger}(\lambda\delta_{ij}-\varrho_{ij})\phi_{j}}. (103)

Then the two point function ϕiϕj¯\overline{\phi_{i}\phi_{j}^{\dagger}} is equal to the resolvent

ϕiϕj¯=(λϱ)ij1.\displaystyle\overline{\phi_{i}\phi_{j}^{\dagger}}=(\lambda-\varrho)^{-1}_{ij}. (104)

We can rewrite the integral as

𝑑ϕ𝑑ϕeϕ(λϱ)ϕ\displaystyle\int d\phi^{\dagger}d\phi e^{-\phi^{\dagger}(\lambda-\varrho)\phi} =𝑑ϕ𝑑ϕ𝑑Gijδ(Gijϕiϕj)eTr(λϱ)G\displaystyle=\int d\phi^{\dagger}d\phi dG_{ij}\delta(G_{ij}-\phi_{i}\phi_{j}^{\dagger})e^{-\operatorname{Tr}(\lambda-\varrho)G} (105)
=𝑑ϕ𝑑ϕ𝑑G𝑑ΣeΣji(Gijϕiϕj)Tr(λϱ)G\displaystyle=\int d\phi^{\dagger}d\phi dGd\Sigma e^{\Sigma_{ji}(G_{ij}-\phi_{i}\phi_{j}^{\dagger})-\operatorname{Tr}(\lambda-\varrho)G}
=𝑑G𝑑ΣeTrΣGTrlogΣTr(λϱ)G.\displaystyle=\int dGd\Sigma e^{\operatorname{Tr}\Sigma G-\operatorname{Tr}\log\Sigma-\operatorname{Tr}(\lambda-\varrho)G}.

The density matrix ϱij\varrho_{ij} is given by

ϱij=CiACjKZ1=CiA^Cj,A^=AKZ1.\displaystyle\varrho_{ij}=\frac{C_{i}^{\dagger}AC_{j}}{KZ_{1}}=C_{i}^{\dagger}\widehat{A}C_{j},\qquad\widehat{A}=\frac{A}{KZ_{1}}. (106)

After integrating out C,CC^{\dagger},C we have

𝑑C𝑑CeTrCC+TrϱG=eTrlog(1A^G),\displaystyle\int dC^{\dagger}dCe^{-\operatorname{Tr}C^{\dagger}C+\operatorname{Tr}\varrho G}=e^{-\operatorname{Tr}\log(1-\widehat{A}G)}, (107)

where Trlog(1A^G)\operatorname{Tr}\log(1-\widehat{A}G) should be understood as the trace of both color and flavor indices. Then (105) becomes 𝑑G𝑑ΣeI\int dGd\Sigma e^{-I} where the action II is given by

I=TrΣG+TrlogΣ+λTrG+Trlog(1A^G).\displaystyle I=-\operatorname{Tr}\Sigma G+\operatorname{Tr}\log\Sigma+\lambda\operatorname{Tr}G+\operatorname{Tr}\log(1-\widehat{A}G). (108)

In the planar approximation, the GG- and Σ\Sigma-integrals can be evaluated by the saddle point approximation. The saddle point equations read

IΣij\displaystyle\frac{\partial I}{\partial\Sigma_{ij}} =Gij+(Σ1)ij=0,\displaystyle=-G_{ij}+(\Sigma^{-1})_{ij}=0, (109)
IGij\displaystyle\frac{\partial I}{\partial G_{ij}} =Σij+λδijδijTrA^1A^G=0.\displaystyle=-\Sigma_{ij}+\lambda\delta_{ij}-\delta_{ij}\operatorname{Tr}\frac{\widehat{A}}{1-\widehat{A}G}=0.

Multiplying the second equation of (109) by GijG_{ij} and summing over i,ji,j, we find

K+λTrGTrGTrA^1A^G=0.\displaystyle-K+\lambda\operatorname{Tr}G-\operatorname{Tr}G\operatorname{Tr}\frac{\widehat{A}}{1-\widehat{A}G}=0. (110)

This is rewritten as

λTrG\displaystyle\lambda\operatorname{Tr}G =K+TrGTrA^1A^G\displaystyle=K+\operatorname{Tr}G\operatorname{Tr}\frac{\widehat{A}}{1-\widehat{A}G} (111)
=K+TrGn=1TrA^nTrGn1.\displaystyle=K+\operatorname{Tr}G\sum_{n=1}^{\infty}\operatorname{Tr}\widehat{A}^{n}\operatorname{Tr}G^{n-1}.

In the planar approximation we have

TrG\displaystyle\operatorname{Tr}G iϕiϕi¯=i(λϱ)ii1=R(λ).\displaystyle\approx\sum_{i}\overline{\phi_{i}\phi_{i}^{{\dagger}}}=\sum_{i}(\lambda-\varrho)^{-1}_{ii}=R(\lambda). (112)

We also find

TrGn\displaystyle\operatorname{Tr}G^{n} =ϕi1ϕi2ϕi2ϕi3ϕinϕi1¯\displaystyle=\overline{\phi_{i_{1}}\phi^{\dagger}_{i_{2}}\phi_{i_{2}}\phi^{\dagger}_{i_{3}}\cdots\phi_{i_{n}}\phi^{\dagger}_{i_{1}}} (113)
ϕi1ϕi1¯ϕi2ϕi2¯ϕinϕin¯\displaystyle\approx\overline{\phi^{\dagger}_{i_{1}}\phi_{i_{1}}}\cdot\overline{\phi^{\dagger}_{i_{2}}\phi_{i_{2}}}\cdots\overline{\phi_{i_{n}}^{\dagger}\phi_{i_{n}}}
=(TrG)n=R(λ)n.\displaystyle=(\operatorname{Tr}G)^{n}=R(\lambda)^{n}.

Finally, (111) becomes

λR(λ)=K+n=1TrA^nR(λ)n.\displaystyle\lambda R(\lambda)=K+\sum_{n=1}^{\infty}\operatorname{Tr}\widehat{A}^{n}R(\lambda)^{n}. (114)

Thus we have re-derived the Schwinger-Dyson equation (22), which was originally derived by means of diagrams in Penington:2019kki . The above saddle point method can be generalized to the Grassmann-odd integral (i.e. to the case of FZZT branes).

References