Holographic spacetime, black holes and quantum error correcting codes: A review

Tanay Kibe    Prabha Mandayam    Ayan Mukhopadhyay
Center for Quantum Information Theory of Matter and Spacetime, and Center for Strings, Gravitation and Cosmology, Department of Physics, Indian Institute of Technology Madras, Chennai 600036, India
tanayk@smail.iitm.ac.in, prabhamd@physics.iitm.ac.in, ayan@physics.iitm.ac.in
(Received: date / Revised version: date)
Abstract

This article reviews the progress in our understanding of the reconstruction of the bulk spacetime in the holographic correspondence from the dual field theory including an account of how these developments have led to the reproduction of the Page curve of the Hawking radiation from black holes. We review quantum error correction and relevant recovery maps with toy examples based on tensor networks, and discuss how it provides the desired framework for bulk reconstruction in which apparent inconsistencies with properties of the operator algebra in the dual field theory are naturally resolved. The importance of understanding the modular flow in the dual field theory has been emphasized. We discuss how the state-dependence of reconstruction of black hole microstates can be formulated in the framework of quantum error correction with inputs from extremal surfaces along with a quantification of the complexity of encoding of bulk operators. Finally, we motivate and discuss a class of tractable microstate models of black holes which can illuminate how the black hole complementarity principle can emerge operationally without encountering information paradoxes, and provide new insights into generation of desirable features of encoding into the Hawking radiation.

journal: Eur. Jour. of Phys. C

1 Introduction

The AdS/CFT correspondence Maldacena:1997re ; Gubser:1998bc ; Witten:1998qj is the most well understood example of the holographic emergence of spacetime and gravity. The heuristic reasoning for the holographic principle of gravity is simply that if we stuff in enough matter in a box, then eventually it will collapse to form a black hole whose maximum possible size would be that of the box Susskind:1994vu . The maximal entropy of a theory of gravity inside the box would then be the Bekenstein-Hawking entropy Bekenstein:1973ur ; Bardeen:1973gs ; Bekenstein:1974ax ; Bekenstein_1981 of the black hole whose horizon is of the size of the box, which explicitly is A/4G𝐴4𝐺A/4G, i.e. the quarter of the area A𝐴A of the horizon measured in Planck units (we set =c=1Planck-constant-over-2-pi𝑐1\hbar=c=1). This heuristic argument relies on a semi-classical description of gravity and should be approximately correct if the size of the box is very large so that gravity is weak even at the black hole horizon when the black hole is of the same size as the box. More precise versions of this argument incorporating covariance under diffeomorphisms and inputs from quantum information theory Bousso:1999cb ; Bousso:1999xy ; Casini:2008cr have been instrumental in providing a concrete ground for the holographic principle which states that a quantum theory of gravity in a spacetime with appropriate asymptotic boundary conditions can be described in terms of a (non-gravitational) quantum many-body system living at the boundary. The AdS/CFT correspondence is a concrete instance of a holographic duality between quantum (super)gravity with asymptotically anti-de Sitter (AdS𝐴𝑑𝑆AdS) boundary conditions (i.e. with constant negative curvature near the boundary) and a (super)conformal gauge theory.

The origin of the AdS/CFT correspondence is from the description of D-branes in open and closed string theory, and is a consequence of open-closed string duality Maldacena:1997re (see Pinaki for a very accessible account). In closed string theory, coincident Dp𝐷𝑝Dp branes are solitonic solutions of ten (or eleven) dimensional supergravity with AdSp+2×X𝐴𝑑subscript𝑆𝑝2𝑋AdS_{p+2}\times X throats, where X𝑋X is a compact space of 8p8𝑝8-p (or 9p9𝑝9-p) dimensions. In open string theory, these coincident branes are p+1𝑝1p+1 spacetime dimensional defects (extended over p𝑝p spatial dimensions), whose low energy descriptions are given by non-Abelian gauge theories living on the worldvolume. A decoupling limit (see Fig. 1) isolates the gauge theory from the remaining stringy degrees of freedom in the open string description, whereas in the closed string description, the excitations living in the AdSp+2×X𝐴𝑑subscript𝑆𝑝2𝑋AdS_{p+2}\times X near-horizon geometry decouples from those in the remaining spacetime. It follows then that the closed string theory (quantum gravity) in AdSp+2×X𝐴𝑑subscript𝑆𝑝2𝑋AdS_{p+2}\times X can be described by a precise p+1𝑝1p+1-dimensional gauge theory111The compact space X𝑋X is related to the global symmetries of the gauge theory. One can obtain more examples of such holographic (a.k.a. gauge/gravity) duality in various dimensions via such string-theoretic setups including those where the gauge theories are non-conformal (and the asymptotic boundary conditions of gravity are non-AdS). We also obtain a precise mapping between the gauge coupling gYMsubscript𝑔𝑌𝑀g_{YM} and the rank of the gauge group (N𝑁N) with the parameters specifying the boundary conditions of the quantum gravity theory, e.g. L𝐿L, the asymptotic curvature radius of AdS𝐴𝑑𝑆AdS and the size of the internal manifold X𝑋X. The AdS/CFT correspondence and its generalizations are often called gauge/gravity dualities Aharony:1999ti .

Refer to caption
(a) A decoupling limit isolates the massless (red) open string excitations in presence of a stack of D-branes. These can be described exactly by a non-Abelian gauge theory.
Refer to caption
(b) The same decoupling limit isolates the closed string excitations of the near-horizon throat geometry in the supergravity solution describing the same stack of D-branes in closed string theory due to an infinite redshift factor. The exact description in this limit is thus the closed string theory in the throat which is asymptotically AdS5×X𝐴𝑑subscript𝑆5𝑋AdS_{5}\times X.
Figure 1: An illustration of how the AdS/CFT correspondence emerges from two dual descriptions of coincident D-branes in open and closed string theory. These figures are from Mateos:2007ay .

The most remarkable aspect of these dualities follow from two features of the holographic dictionary. Firstly, the ’t Hooft coupling λ=gYM2N𝜆superscriptsubscript𝑔𝑌𝑀2𝑁\lambda=g_{YM}^{2}N is related to a positive power of L/ls𝐿subscript𝑙𝑠L/l_{s} (lssubscript𝑙𝑠l_{s} is the string length) that controls the corrections to a two-derivative Einsteinian gravity theory arising from the finite length of the string. Furthermore, N𝑁N, the rank of the gauge group, is related to a positive power of L/lp𝐿subscript𝑙𝑝L/l_{p} (lpsubscript𝑙𝑝l_{p} is the Planck length) that controls the quantum corrections to classical gravity. As a result, when both N𝑁N and λ𝜆\lambda are large, both lssubscript𝑙𝑠l_{s} and lpsubscript𝑙𝑝l_{p} are small in units where the asymptotic curvature radius L𝐿L is set to unity, so that both stringy and quantum effects are suppressed. Thus the dual description is just a classical gravity theory to a very good approximation. The duality therefore implies that a many-body system which evades a quasi-particle description can be described by a classical Einsteinian theory of gravity in one higher dimension with a negative cosmological constant and minimally coupled to a few fields. This has led to an enormous impact on our understanding of the collective description of many strongly interacting systems, including strongly correlated quantum materials Hartnoll:2016apf , non-perturbative dynamics of QCD Kim:2012ey ; Rebhan:2014rxa and its various phases such as the quark-gluon plasma Casalderrey-Solana:2011dxg .

Over more than two decades, the correspondence has also been subjected to stringent tests, where non-trivial dual quantities such as the anomalous dimensions of single-trace gauge-invariant operators and the spectrum of strings in anti-de Sitter space have been matched using techniques like integrability Beisert:2010jr and localization Zarembo:2016bbk . Recently a derivation of the correspondence has been achieved when the dual gauge theory is free (zero ’t Hooft coupling) while the string worldsheet sees a quantum spacetime and gets localized at the boundary Eberhardt:2019ywk ; Gaberdiel:2021qbb .

The fundamental aspects of how the bulk spacetime and its gravitational dynamics emerge from the dual gauge theory are still shrouded in many mysteries. Nevertheless, there has been remarkable progress in this direction via the tools of quantum information theory. A path breaking proposal by Ryu and Takanayagi RT and its further refinements HRT ; EngelhardtWall that an appropriate codimension two bulk extremal surface anchored to the boundary R𝑅\partial R of a boundary spatial subregion R𝑅R captures the entanglement entropy of that region R𝑅R in the dual field theory, have been at the heart of these developments. To be specific, we need to consider the causal domain of dependence DRsubscript𝐷𝑅D_{R} of a region R𝑅R as the set of points where the values of measurements can necessarily be influenced by or influence the data on R𝑅R. Then the bulk operators in the entanglement wedge, which is the causal domain of dependence of (any) Cauchy slice bounded by the bulk extremal surface anchored to R𝑅\partial R and the subregion R𝑅R at the boundary, can be decoded from the algebra of operators in the dual CFT in DRsubscript𝐷𝑅D_{R} as illustrated in Fig. 9 Czech_2012 ; Wall:2012uf ; FLM ; Headrick_2014 ; Jafferis_2016 ; Faulkner_2017 .

Subsequently, quantum information theory has played a fundamental role in understanding how this entanglement wedge of the emergent spacetime can be reconstructed in the dual conformal field theory (CFT) without encountering inconsistencies. It has been shown that the correct framework which achieves bulk reconstruction in a consistent way can be obtained by reformulating the AdS/CFT correspondence as a quantum error correcting code in which the bulk spacetime is encoded in a redundant way in the Hilbert space of the CFT. Furthermore, the encoding is protected against deletion errors, i.e. allowing for the entanglement wedge to be (approximately) reconstructed from its corresponding boundary subregion even after the complement of the latter is traced out Almheiri_2015 ; Harlow:2016vwg ; dong_2016_2 ; cotler2019_univR .

The connection between the holographic principle of gravity and quantum information theory is currently a topic of fundamental interest to researchers in diverse fields. In fact, this interdisciplinary area of research has been instrumental in developing new perspectives in quantum error correction itself, and has produced novel connections between quantum fields (many-body systems) and quantum information theory as well. One example of such a connection is the postulate of the quantum null-energy condition which states that the expectation value of a null projection of the energy-momentum tensor is bounded from below by a specific null variation of the entanglement entropy Bousso_2016 . This postulate has not only been proven in holographic field theories Koeller_2016 , but also in generic two-dimensional CFTs Balakrishnan:2017bjg and free field theories Bousso_2016_2 ; Malik:2019dpg , and is also expected to hold generally. Furthermore, such developments have led to new understanding of connections between entanglement and the renormalization group (RG) flow especially with respect to the existence of quantities which evolve monotonically under the flow Freedman:1999gp ; Myers:2010xs ; Casini:2012ei .

The most fundamental test of our understanding of the AdS/CFT correspondence is whether we can find explicit mechanisms for the resolution of black hole information paradoxes which are further refinements of Hawking’s original result that a semi-classical black hole should lose its mass to thermal (Hawking) radiation violating unitarity Hawking:1975vcx ; Hawking:1976ra . The understanding of the AdS/CFT correspondence in the information theory framework has driven remarkable progress in this frontier as well. In particular, it has been shown that the AdS/CFT correspondence itself leads us to the correct way to compute the fine grained entanglement entropy of the Hawking radiation222Following Ghosh:2021axl , we should call it the ”not-so-fine-grained entropy” actually. It is assumed that some averaging has been done due to which an approximate notion of factorization of the Hilbert space into black hole interior and radiation is valid for observables of an effective field theory. Only in such an operational context, a Page curve can be defined suitably. In this review, we will discuss a class of microstate models in Section 5.3.2 which illustrates some aspects of how such factorizations can emerge. In Ghosh:2021axl , there is a discussion on this issue within the semi-classical approximation in the context of braneworld cosmology.. Surprisingly, these computations can be done in the semi-classical evaporating black hole geometry itself and produce results which are consistent with unitarity. In setups where the holographic system is connected to a bath which collects the Hawking quanta of an evaporating black hole, the bath develops an entanglement wedge, called the island which contains portions of the interior of the black hole, once the black hole is past the Page time (approximately the time when the black hole and the Hawking radiation have the same number of degrees of freedom). The inclusion of this spatially disconnected island leads to reproduction the Page curve Page_1993 ; Page_2013 for the time-dependence of the von Neumann entropy of the Hawking radiation in consistency with unitarity without invalidating the effective semiclassical description of bulk physics AEMM ; PeningtonQES ; AlmheiriQES ; Penington:2019kki ; Almheiri2020Islands . The further understanding of how information of the black hole interior is encoded without encountering fundamental inconsistencies is probably the most exciting topic at the intersection of quantum information theory and gravity.

The present review is aimed to provide an accessible account for researchers in diverse fields to follow the developments connecting the holographic emergence of spacetime and gravity with quantum information theory. We also give a special emphasis on recent developments in connection with black holes. Our account is somewhat complementary to existing reviews in literature, and is also self-contained. As instances of focused reviews on subtopics covered here, we would especially like to mention Rangamani:2016dms which reviews the holographic entanglement entropy proposal, Mathur:2009hf ; Harlow:2014yka ; Raju:2020smc which review the black hole information puzzles and their possible resolutions, harlow2018tasi ; Jahn:2021uqr ; Chen:2021lnq which review aspects of bulk reconstruction, and Almheiri:2020cfm (see mahajan for a very accessible summary) which reviews recent progress in reproduction of the Page curve (the time-dependence of the entanglement entropy of Hawking radiation) from the AdS/CFT correspondence. We also update the content of these reviews, and describe the central concepts and proposals along with essences of their derivations or arguments supporting them in a sufficiently detailed manner. Furthermore, we present perspectives on the open questions and some promising directions for research in the near future.

The plan of the review is as follows. In section 2, we introduce the Ryu-Takanayagi surface and its generalization the quantum extremal surface which is the key to compute the entanglement entropy of a subregion in the dual field theory. As mentioned before, this is at the heart of the connection between quantum information theory and the holographic correspondence. We also review the proof of these proposals for bulk extremal surfaces at leading and subleading orders. We furthermore discuss the consistency checks for the proposal for the quantum extremal surface. We especially outline the proof that the entanglement wedge contains the bulk causal wedge which is the key to the understanding of the non-triviality of bulk reconstruction. Furthermore, we discuss how the holographic presciptions for computing entanglement entropy reproduce entanglement inequalities (especially the strong subadditivity) and the alternative maximin construction.

In section 3, we introduce the entanglement wedge reconstruction hypothesis and the key postulate of equality of bulk and boundary relative entropies in the bulk semi-classical approximation. We then discuss various implications of the latter postulate especially how it implies the emergence of gravitational field equations at linearized order, and furthermore how the bulk canonical energy gets connected to Fisher information at the boundary. We also discuss progress in understanding of the emergence of bulk from modular flow at the boundary, and the basic reasons for reformulating bulk reconstruction as a quantum error correcting code. Afterwards, we introduce the appearance of islands which are disconnected portions of the entanglement wedge especially in the context of double holography, and introduce and justify the island rule for computing the entanglement entropy of a subregion of a bath in contact with a holographic system described by a semiclassical evaporating black hole while sketching how this reproduces the Page curve of Hawking radiation.

In section 4, we review quantum error correction especially in relation to operator algebras, and discuss how this framework together with the postulate of equality of bulk and boundary relative entropies imply reconstruction of local bulk operators in the entanglement wedge in terms of the operators of the dual field theory in the region of interest. We discuss how apparent inconsistencies of bulk reconstruction are mitigated by formulating the bulk reconstruction in AdS/CFT correspondence as a quantum error correcting code that protects against deletion of complementary boundary subregions. Toy models of AdS/CFT correspondence based on tensor networks which provide examples of perfect recovery maps for the entanglement wedge in the dual subregion of the field theory are then presented. We focus particularly on the Petz map, and discuss how a specific variation could provide the desired approximate recovery even at sub-leading order, and relate to the modular flow at the boundary. We furthermore discuss the connection between the bulk radial coordinate and renormalization group flow in this context with a perspective on issues which may spur further universal understanding of the holographic correspondence.

In section 5, we review the exciting recent progress in the reproduction of the Page curve from Hawking radiation. In particular, we focus on how replica wormhole saddles imply the quantum extremal surfaces and islands responsible for restoring behavior of the Page curve that is consistent with unitarity although these wormholes provide an intrinsically averaged description of the black hole microstates. We then discuss the state-dependence of the encoding of the interior in the framework of universal (approximate) subsystem recovery and also how the Python’s lunch mechanism generates exponential complexity of the state-dependent encoding.

Furthermore, we present a perspective on issues that are crucial to fully understand how the black hole complementarity principle can emerge operationally without encountering information paradoxes such as the AMPS paradox while emphasizing the need for a complex encoding of the interior. We motivate the need for tractable microstate models of black holes which can demonstrate the realization of all desirable features of the encoding in the Hawking radiation especially information mirroring (with decoding possible without full knowledge of interior) and exponentially complex encoding of the black hole interior excitations simultaneously. Furthermore, it should explain the origin of self-averaging in real time to be consistent with the implications of replica wormholes. We proceed to discuss a class of tractable microstate models explicitly, their promising results in this direction along with implications and some open questions.

In section 6, we conclude with a discussion on some of the topics of significance which are not covered in this review and some promising directions for further research.

2 Quantum extremal surfaces

2.1 Partial proof of the quantum extremal surface proposal

The entanglement entropy of a boundary subregion provides the most fundamental link between holography and quantum information.

Ryu and Takayanagi (RT) RT were the first to describe how entanglement entropy could be computed holographically in a static semi-classical spacetime (dual to a large N𝑁N quantum field theory). Then Hubeny, Rangamani and Takayanagi (HRT) HRT extended this idea to time dependent geometries. The RT/HRT formula states that the entanglement entropy of a subregion R of the boundary theory is proportional to the area of the classical, bulk co-dimension two 333If N𝑁N is a submanifold of M𝑀M then the codimension of N𝑁N in M𝑀M is defined to be: codim(N)=dim(M)dim(N)𝑐𝑜𝑑𝑖𝑚𝑁𝑑𝑖𝑚𝑀𝑑𝑖𝑚𝑁codim(N)=dim(M)-dim(N). extremal surface that is anchored to the boundary of R and is homologous to R (see Fig. 2). The RT proposal was shown to be correct at the leading order in Planck-constant-over-2-pi\hbar by Lewkowycz and Maldacena (LM)LM . A generalization of this proof for the HRT proposal using a bulk version of the Schwinger-Keldysh contour was described in Dong_2016 . The RT/HRT prescription has led to a much simpler proof of the strong sub-additivity of entanglement entropy Headrick:2007km ; Wall:2012uf and also to new inequalities that holographic entanglement entropy should satisfy Hayden:2011ag ; Bao:2015bfa ; Hubeny:2018ijt ; Hubeny:2018trv . The next to leading order correction to the holographic entanglement entropy for semi-classical static situations was computed by Faulkner, Lewkowycz and Maldacena (FLM) FLM . They found that the entanglement entropy correct to 𝒪(0)𝒪superscriptPlanck-constant-over-2-pi0\mathcal{O}(\hbar^{0}) is given by

Refer to caption
Figure 2: The red region is the boundary sub-region of interest (R). The classical extremal surface anchored to the boundary of R is shown.
S(R)=𝒜(XR)4G+Sent-bulk=Sgen(XR),𝑆𝑅𝒜subscript𝑋𝑅4𝐺Planck-constant-over-2-pisubscript𝑆ent-bulksubscript𝑆gensubscript𝑋𝑅S(R)=\frac{\mathcal{A}(X_{R})}{4G\hbar}+S_{\text{ent-bulk}}=S_{\text{gen}}(X_{R}), (1)

where 𝒜(XR)𝒜subscript𝑋𝑅\mathcal{A}(X_{R}) is the area of the classical bulk extremal HRT surface (XRsubscript𝑋𝑅X_{R}) and the leading order quantum corrections are given by the bulk entanglement entropy (Sent-bulksubscript𝑆ent-bulkS_{\text{ent-bulk}}) which is as follows. As shown in Fig. 3, the classical extremal surface divides the bulk into two sub-regions Rbsubscript𝑅𝑏R_{b} and R¯bsubscript¯𝑅𝑏\bar{R}_{b}; Sent-bulksubscript𝑆ent-bulkS_{\text{ent-bulk}} is the entanglement entropy of the reduced density matrix (in the bulk effective field theory) on the bulk sub-region Rbsubscript𝑅𝑏R_{b} that is connected to the boundary region R𝑅R. The bulk quantum fields are assumed to be described by an effective field theory (EFT) living on a fixed background and the entanglement entropy of the bulk sub-region connected to the boundary region R𝑅R is computed using standard quantum field theory techniques. The quantity defined in Eq. (1) is called the generalized entropy. The quantity Sent-bulksubscript𝑆ent-bulkS_{\text{ent-bulk}} suffers from divergences, which can be absorbed into the renormalization of the Newton’s constant leading to a well defined generalized entropy Srednicki:1993im ; Susskind:1994sm ; Kabat:1995eq ; Larsen:1995ax ; Jacobson:1994iw . It is not clear if the LM and FLM proofs (reviewed below) can be extended to include higher order corrections involving multiple loops of quantized gravitons. However, the Engelhardt-Wall proposal is expected to work for all orders in Planck-constant-over-2-pi\hbar in which we consider the full quantum theory of the bulk matter on a fixed (but backreacted) semiclassical gravitational background as discussed below.

Engelhardt and Wall (EW) EngelhardtWall conjectured that the holographic entanglement entropy of a boundary sub-region R𝑅R is given by the generalized entropy of the quantum extremal surface (QES) χRsubscript𝜒𝑅\chi_{R} anchored to (R𝑅\partial R), so that

S(R)=Sgen(χR),𝑆𝑅subscript𝑆gensubscript𝜒𝑅S(R)=S_{\text{gen}}(\chi_{R}), (2)

where the quantum extremal surface is the surface that extremizes the generalized entropy and is homologous to R𝑅R. If there are multiple such extremal surfaces, the one with the smallest generalized entropy satisfying the homology constraint is picked.444Beyond the semiclassical gravity limit, the area of the quantum extremal surface should be promoted to an area operator A^^𝐴\hat{A} which satisfies the identity δA^δXaρ=0=ρδA^δXa,𝛿^𝐴𝛿superscript𝑋𝑎𝜌0𝜌𝛿^𝐴𝛿superscript𝑋𝑎\frac{\delta\hat{A}}{\delta X^{a}}\rho=0=\rho\frac{\delta\hat{A}}{\delta X^{a}}, where ρ𝜌\rho is a state of the Hilbert space of the bulk matter theory as discussed in EngelhardtWall . Furthermore, we will need to consider A+Sentbulk+countertermsdelimited-⟨⟩𝐴subscript𝑆entbulkdelimited-⟨⟩counterterms\langle A\rangle+S_{\rm ent-bulk}+\langle{\rm counterterms}\rangle to define the holographic entanglement entropy (where counterterms remove the ultraviolet divergences of the area operator). The validity of such an approach has been examined in Belin:2018juv ; Belin:2019mlt ; Belin:2021htw perturbatively in the 1/N1𝑁1/N expansion. This QES proposal is different from the FLM formula, which is the generalized entropy of the classical RT/HRT surface.

Refer to caption
Figure 3: The red region is the boundary sub-region of interest. The magenta extremal surface divides the bulk into two regions. Rbsubscript𝑅𝑏R_{b} is the bulk sub-region connected to the boundary region of interest.

The QES proposal hasn’t been proven, however it is believed to be correct since it satisfies some non trivial consistency checks which we will review in section 2.2. This proposal has been checked order by order in Planck-constant-over-2-pi\hbar for a few sub-leading corrections by comparing computations in the bulk with those in the dual boundary CFT Belin:2018juv ; Belin:2019mlt ; Belin:2021htw . In the rest of this review, we will use units where =1Planck-constant-over-2-pi1\hbar=1 unless explicitly mentioned.

Before describing the checks on the QES proposal, we will first review the classical argument from LM that proves the RT formula. The entanglement entropy of a boundary sub-region R𝑅R can be computed using the replica trick. This consists of going to Euclidean time and considering an angular direction in the boundary field theory with origin at the boundary of R𝑅R. This is labelled by τ𝜏\tau, with τ=τ+2π𝜏𝜏2𝜋\tau=\tau+2\pi. The boundary quantum field theory (QFT) is then considered in a sequence of spaces (M~nsubscript~𝑀𝑛\widetilde{M}_{n}) with τ=τ+2πn𝜏𝜏2𝜋𝑛\tau=\tau+2\pi n, for positive integers n. This sequence is holographically dual to a sequence of bulk geometries labelled by B~nsubscript~𝐵𝑛\widetilde{B}_{n} with the asymptotic boundary M~nsubscript~𝑀𝑛\widetilde{M}_{n}. One then computes the Rènyi entropy (Snsubscript𝑆𝑛S_{n}) as follows:

Sn=11n(lnZ[M~n]nlnZ[M~1]).subscript𝑆𝑛11𝑛𝑍delimited-[]subscript~𝑀𝑛𝑛𝑍delimited-[]subscript~𝑀1S_{n}=\frac{1}{1-n}(\ln Z[\widetilde{M}_{n}]-n\ln Z[\widetilde{M}_{1}]). (3)

Analytic continuation to non-integer n𝑛n is well defined due to the Carlson theorem carlson . The n1𝑛1n\to 1 limit gives the von Neumann entropy. Here Z[M~n]𝑍delimited-[]subscript~𝑀𝑛Z[\widetilde{M}_{n}] and Z[M~1]𝑍delimited-[]subscript~𝑀1Z[\widetilde{M}_{1}] are the QFT partition functions for the spacetime Mn~~subscript𝑀𝑛\widetilde{M_{n}} and the original spacetime M1subscript𝑀1M_{1}. The holographic dictionary in the large N𝑁N limit tells us the following:

Z[M~n]=eI[B~n],𝑍delimited-[]subscript~𝑀𝑛superscript𝑒𝐼delimited-[]subscript~𝐵𝑛Z[\widetilde{M}_{n}]=e^{-I[\widetilde{B}_{n}]}, (4)

where I[B~n]𝐼delimited-[]subscript~𝐵𝑛I[\widetilde{B}_{n}] is the on-shell classical bulk action for the bulk geometry B~nsubscript~𝐵𝑛\widetilde{B}_{n}. The bulk geometries B~nsubscript~𝐵𝑛\widetilde{B}_{n} have a nsubscript𝑛\mathbb{Z}_{n} symmetry that corresponds to cyclic permutations of the n replicas. Taking a quotient with this, one can define Bn=B~n/nsubscript𝐵𝑛subscript~𝐵𝑛subscript𝑛B_{n}=\widetilde{B}_{n}/\mathbb{Z}_{n}. Due to the quotient, these bulk geometries have the same boundary conditions as the original geometry (B~1subscript~𝐵1\widetilde{B}_{1}), that is τ=τ+2π𝜏𝜏2𝜋\tau=\tau+2\pi. These geometries typically have a conical defect with opening angle 2πn2𝜋𝑛\frac{2\pi}{n} at the fixed points of the nsubscript𝑛\mathbb{Z}_{n} symmetry. The classical bulk action is a τ𝜏\tau integral of a local Lagrangian density, therefore it follows that I[B~n]=nI[Bn]𝐼delimited-[]subscript~𝐵𝑛𝑛𝐼delimited-[]subscript𝐵𝑛I[\widetilde{B}_{n}]=nI[B_{n}]. Thus the Rènyi entropy can be written as follows:

Sn=n1n(I[Bn]I[B1]).subscript𝑆𝑛𝑛1𝑛𝐼delimited-[]subscript𝐵𝑛𝐼delimited-[]subscript𝐵1S_{n}=\frac{n}{1-n}(I[B_{n}]-I[B_{1}]). (5)

Note that when evaluating I[Bn]𝐼delimited-[]subscript𝐵𝑛I[B_{n}] one excludes any contributions from the conical singularity. We can now analytically continue to non-integer n and take the n1𝑛1n\to 1 limit. The von Neumann entropy is then:

S=nI[Bn]𝑆subscript𝑛𝐼delimited-[]subscript𝐵𝑛S=-\partial_{n}I[B_{n}] (6)

Varying n𝑛n corresponds to changing the opening angle of the conical defect. The metric and other fields also have to change due to this change in n𝑛n. However since the geometry Bnsubscript𝐵𝑛B_{n} is a solution of the equations of motion, the first order variations of the bulk action away from these solutions should vanish. Thus the only change in the action comes from a boundary term (at the conical singularity). Therefore, the von Neumann entropy is essentially a boundary term at the conical singularity, which is a co-dimension 2 hypersurface in the bulk. This boundary term was calculated in LM and was shown to reproduce the RT formula.

The above argument was extended by FLM FLM to include quantum corrections (to 𝒪(0)𝒪superscriptPlanck-constant-over-2-pi0\mathcal{O}(\hbar^{0})) by considering quantum fluctuations in the bulk EFT. The partition function of bulk quantum fields is then given by the following:

Zbulk(n)=Tr[ρnn],superscriptsubscript𝑍𝑏𝑢𝑙𝑘𝑛𝑇𝑟delimited-[]subscriptsuperscript𝜌𝑛𝑛Z_{bulk}^{(n)}=Tr[\rho^{n}_{n}], (7)

where ρnsubscript𝜌𝑛\rho_{n} is a state of the bulk quantum fields in the bulk geometry Bnsubscript𝐵𝑛B_{n} . The Rènyi entropy can now be written as follows:

Sn=11n(lnTr[ρnn]nlnTr[ρ1]).subscript𝑆𝑛11𝑛𝑇𝑟delimited-[]subscriptsuperscript𝜌𝑛𝑛𝑛𝑇𝑟delimited-[]subscript𝜌1S_{n}=\frac{1}{1-n}(\ln Tr[\rho^{n}_{n}]-n\ln Tr[\rho_{1}]). (8)

The von Neumann entropy is then:

S=n(lnTr[ρnn]nlnTr[ρ1])n=1.𝑆subscript𝑛subscript𝑇𝑟delimited-[]subscriptsuperscript𝜌𝑛𝑛𝑛𝑇𝑟delimited-[]subscript𝜌1𝑛1S=-\partial_{n}(\ln Tr[\rho^{n}_{n}]-n\ln Tr[\rho_{1}])_{n=1}. (9)

We can add and subtract the term n(lnTr[ρ1n])subscript𝑛𝑇𝑟delimited-[]superscriptsubscript𝜌1𝑛-\partial_{n}(\ln Tr[\rho_{1}^{n}]) to Eq. (9) and after some algebra we get:

S=Sent-bulk+Sarea,𝑆subscript𝑆ent-bulksubscript𝑆areaS=S_{\text{ent-bulk}}+S_{\text{area}}, (10)

with

Sent-bulk=n(lnTr[ρ1n]nlnTr[ρ1])n=1subscript𝑆ent-bulksubscript𝑛subscript𝑇𝑟delimited-[]superscriptsubscript𝜌1𝑛𝑛𝑇𝑟delimited-[]subscript𝜌1𝑛1S_{\text{ent-bulk}}=-\partial_{n}(\ln Tr[\rho_{1}^{n}]-n\ln Tr[\rho_{1}])_{n=1} (11)

and

Sarea=Tr[nρn]n=1Tr[ρ1].subscript𝑆area𝑇𝑟subscriptdelimited-[]subscript𝑛subscript𝜌𝑛𝑛1𝑇𝑟delimited-[]subscript𝜌1S_{\text{area}}=-\frac{Tr[\partial_{n}\rho_{n}]_{n=1}}{Tr[\rho_{1}]}. (12)

The Sent-bulksubscript𝑆ent-bulkS_{\text{ent-bulk}} term involves ρ1subscript𝜌1\rho_{1} which is the density matrix of bulk quantum fields in the original geometry. This therefore computes the bulk entanglement entropy. The Sareasubscript𝑆areaS_{\text{area}} term can be expressed as a variation of a local Lagrangian, which for the usual 2 derivative gravity action gives the area term as seen before for the classical case. This concludes a heuristic review of the arguments of FLM for the quantum correction to the RT formula.

We will now define the causal wedge and the entanglement wedge. These are important for bulk reconstruction, which is the major focus of this review. For a boundary sub-region R𝑅R, the boundary domain of dependence of R𝑅R, labelled by DRsubscript𝐷𝑅D_{R} is defined to be the set of all boundary points such that any in-extensible timelike curve that passes through any point in DRsubscript𝐷𝑅D_{R} necessarily intersects R𝑅R (i.e. D(R)𝐷𝑅D(R) is the set of points where the values of measurements can necessarily be influenced by or influence the data on R𝑅R). The bulk casual wedge (WRsubscript𝑊𝑅W_{R}) is then defined to be the intersection of the causal past and future of R𝑅{R}. WR=𝒥+(DR)𝒥(DR)subscript𝑊𝑅superscript𝒥subscript𝐷𝑅superscript𝒥subscript𝐷𝑅W_{R}=\mathcal{J^{+}}(D_{R})\cap\mathcal{J^{-}}(D_{R}) The boundary of WRsubscript𝑊𝑅W_{R} is called the causal surface and is labelled by CRsubscript𝐶𝑅C_{R}. See Fig.4 for a pictorial depiction of these definitions.

Refer to caption
Figure 4: The boundary domain of dependence is coloured purple, the causal wedge is the green bulk region and the causal surface is coloured blue. The quantum extremal surface is marked as χRsubscript𝜒𝑅\chi_{R} and is shown to lie deeper into the bulk than the causal surface. Figure from EngelhardtWall .

Classically it should be possible to reconstruct any bulk operator within WRsubscript𝑊𝑅W_{R} in terms of boundary operators on DRsubscript𝐷𝑅D_{R} since they are in causal contact with each other. Entanglement wedge reconstruction however states that any operator in the entanglement wedge of R𝑅R can be reconstructed from operators in DRsubscript𝐷𝑅D_{R}, where the entanglement wedge is defined to be the bulk domain of dependence of the Cauchy surface that interpolates between R𝑅R and the extremal surface anchored to R𝑅\partial R. This has led to the notion of sub-region duality Bousso:2012mh ; Czech:2012bh ; Bousso:2012sj , which states that sub-regions of the boundary are dual to sub-regions of the bulk. Bulk reconstruction will be described in more detail in section 3. We will now use the definitions from above to review certain checks on the QES proposal.

Sanity checks

:Any proposal for the holographic entanglement entropy that claims to be correct to all orders in Planck-constant-over-2-pi\hbar must pass the following preliminary checks.

  1. 1.

    It must agree with the RT/HRT formula at leading order in Planck-constant-over-2-pi\hbar

  2. 2.

    It must agree with the FLM formula at next to leading order in Planck-constant-over-2-pi\hbar

Engelhardt and Wall EngelhardtWall argued that their proposal indeed passes these two checks. We review these arguments below.

Since the FLM proposal is valid only upto 𝒪(0)𝒪superscriptPlanck-constant-over-2-pi0\mathcal{O}(\hbar^{0}), it is enough to show that the following is true:

Sgen(XR)=Sgen(χR)+𝒪().subscript𝑆𝑔𝑒𝑛subscript𝑋𝑅subscript𝑆𝑔𝑒𝑛subscript𝜒𝑅𝒪Planck-constant-over-2-piS_{gen}(X_{R})=S_{gen}(\chi_{R})+\mathcal{O}(\hbar). (13)

The generalized entropy consists of two terms Sgen(X)=𝒜(X)4G+Sentsubscript𝑆𝑔𝑒𝑛𝑋𝒜𝑋4𝐺Planck-constant-over-2-pisubscript𝑆𝑒𝑛𝑡S_{gen}(X)=\frac{\mathcal{A}(X)}{4G\hbar}+S_{ent} . In a semi-classical bulk the classical and quantum extremal surfaces are expected to be a distance 𝒪()𝒪Planck-constant-over-2-pi\mathcal{O}(\hbar) apart. Thus the entanglement entropies of bulk fields (Sentsubscript𝑆𝑒𝑛𝑡S_{ent}) for the two surfaces are expected to differ only at 𝒪()𝒪Planck-constant-over-2-pi\mathcal{O}(\hbar). That is Sent(XR)Sent(χR)=𝒪()subscript𝑆𝑒𝑛𝑡subscript𝑋𝑅subscript𝑆𝑒𝑛𝑡subscript𝜒𝑅𝒪Planck-constant-over-2-piS_{ent}(X_{R})-S_{ent}(\chi_{R})=\mathcal{O}(\hbar). Now we can look at the area term. Since the classical extremal surface extremizes the area, first order variations of the classical surface do not affect the area. That is since XRsubscript𝑋𝑅X_{R} and χRsubscript𝜒𝑅\chi_{R} are a distance Planck-constant-over-2-pi\hbar apart, the leading order difference in the two areas is at most 2superscriptPlanck-constant-over-2-pi2\hbar^{2}, that is A(XR)A(χR)=𝒪(2)𝐴subscript𝑋𝑅𝐴subscript𝜒𝑅𝒪superscriptPlanck-constant-over-2-pi2A(X_{R})-A(\chi_{R})=\mathcal{O}(\hbar^{2}). This therefore proves that the holographic entanglement entropy proposal agrees with the RT/HRT and FLM formulas at the appropriate orders. The FLM formula and the QES proposal will not agree at higher orders in Planck-constant-over-2-pi\hbar and one can perform computations at higher orders to determine if the QES proposal is correct Belin:2018juv ; Belin:2019mlt ; Belin:2021htw . The QES proposal passes some essential consistency checks and is therefore believed to be correct. These checks will be the focus of the rest of this section.

2.2 The entanglement wedge contains the causal wedge

If the sub-region duality is consistent then the QES must lie deeper in the bulk than causal surfaces. This is an important consistency check that the QES proposal passes. In this section we reason why this consistency condition should hold and review a proof of why the QES proposal passes this check.

Let us assume that the extremal surface can lie closer to the boundary than the causal surface. Let us consider a pure state in the dual theory at the boundary. The entanglement entropy of its reduced density matrices on R𝑅R and its complement (ρRsubscript𝜌𝑅\rho_{R} and ρR¯subscript𝜌¯𝑅\rho_{\bar{R}}) must therefore be the same. This can be reproduced by the QES proposal if XR=XR¯subscript𝑋𝑅subscript𝑋¯𝑅X_{R}=X_{\bar{R}} since the bulk fields are also in a pure state. As shown in Fig. 5 if XRsubscript𝑋𝑅X_{R} lies within the causal wedge WRsubscript𝑊𝑅W_{R} then the region between CRsubscript𝐶𝑅C_{R} and XRsubscript𝑋𝑅X_{R} would be in the entanglement wedge of R¯¯𝑅\bar{R} and can therefore be reconstructed on R¯¯𝑅\bar{R}. However this region is in causal contact with DRsubscript𝐷𝑅D_{R} and therefore can be affected by a signal propagating from DRsubscript𝐷𝑅D_{R}. This leads to a contradiction in the dual boundary theory since R𝑅R and R¯¯𝑅\bar{R} are causally disconnected. This can be avoided only if XRsubscript𝑋𝑅X_{R} lies deeper in the bulk than CRsubscript𝐶𝑅C_{R} and is spacelike to it.

Refer to caption
Figure 5: This is a spatial slice of the geometry. The boundary is the circumference of the circle. R𝑅R (red) is the boundary subregion of interest and R¯¯𝑅\bar{R} is its complement. CRsubscript𝐶𝑅C_{R} (magenta) is the causal surface for R𝑅R and XR=XR¯subscript𝑋𝑅subscript𝑋¯𝑅X_{R}=X_{\bar{R}} (black) is the classical extremal surface, which is shown to lie within the causal wedge of R. The grey region is the entanglement wedge of R¯¯𝑅\bar{R}. Anything in the grey region can be reconstructed from R¯¯𝑅\bar{R}. A signal from R can propagate upto CRsubscript𝐶𝑅C_{R} and therefore into the entanglement wedge of R¯¯𝑅\bar{R}. This violates micro-causality of the CFT.

It was shown in HubenyRangamani that if the classical null energy condition (NEC) holds then the classical extremal surface lies deeper in the bulk than the causal surface. The NEC states that Tkk0subscript𝑇𝑘𝑘0T_{kk}\geq 0 for any future directed null vector kμsuperscript𝑘𝜇k^{\mu}. This can be violated if there is quantum matter in the bulk and the classical extremal surface can therefore be closer to the boundary than CRsubscript𝐶𝑅C_{R} or it could be timelike separated from it. The way to avoid this inconsistency is to use the QES instead of the classical extremal surface. Below we review the argument from EngelhardtWall which shows that the QES χRsubscript𝜒𝑅\chi_{R} lies deeper in the bulk than CRsubscript𝐶𝑅C_{R} and is spacelike to it.

We will first state the generalized second law (GSL). The usual second law of thermodynamics states that the thermodynamic entropy of any closed system is nondecreasing in time. The GSL is a statment about the monotonicity of the generalized entropy Wall_2013 . The generalized entropy is computed on a Cauchy slice (at some ”time”), the GSL then states that the variation of this generalized entropy along any future directed normal to the Cauchy slice is non negative. The generalized entropy can be defined for any causal horizon (H+superscript𝐻H^{+}), which is defined to be the boundary of the past of any future directed timelike or null worldline. Define H=H+Σ𝐻superscript𝐻ΣH=H^{+}\cap\Sigma, where ΣΣ\Sigma is a Cauchy slice and H+superscript𝐻H^{+} is a future causal horizon. Then the GSL states the following EngelhardtWall ; Wall_2013

δSgen(H)δHμkμ0,𝛿subscript𝑆𝑔𝑒𝑛𝐻𝛿superscript𝐻𝜇superscript𝑘𝜇0\frac{\delta S_{gen}(H)}{\delta H^{\mu}}k^{\mu}\geq 0, (14)

where δHμ𝛿superscript𝐻𝜇\delta H^{\mu} is a normal to H and kμsuperscript𝑘𝜇k^{\mu} is any future directed null vector. We will also require the following theorem from Wall_2013 , see also EngelhardtWall .

Theorem 2.1.

Let M and N be co-dimension one null surfaces that split the spacetime into two parts, an interior (Int) and exterior (Ext), where Ext is defined to be the region containing the boundary subregion DRsubscript𝐷𝑅D_{R} which is of interest. Let MExt(N)𝑀𝐸𝑥𝑡𝑁M\cap Ext(N) be empty. Also assume that M and N coincide at some point p and that M and N are smooth in the neighbourhood of p. Let ΣΣ\Sigma be a spatial slice that passes through p. Then there exists a normal (δΣμ𝛿superscriptΣ𝜇\delta\Sigma^{\mu}) to ΣMΣ𝑀\Sigma\cap M in the neighbourhood of p𝑝p such that

δSgen(M)δΣμkμδSgen(N)δΣμkμ,𝛿subscript𝑆𝑔𝑒𝑛𝑀𝛿superscriptΣ𝜇superscript𝑘𝜇𝛿subscript𝑆𝑔𝑒𝑛𝑁𝛿superscriptΣ𝜇superscript𝑘𝜇\frac{\delta S_{gen}(M)}{\delta\Sigma^{\mu}}k^{\mu}\geq\frac{\delta S_{gen}(N)}{\delta\Sigma^{\mu}}k^{\mu}, (15)

where kμsuperscript𝑘𝜇k^{\mu} is a future directed null normal to M𝑀M and N𝑁N.

To use this theorem and show that the QES lies deeper than the causal surface we identify the null splitting surface M𝑀M with the boundary of the entanglement wedge. This can be generated by shooting null rays from χRsubscript𝜒𝑅\chi_{R} towards R𝑅R. Let us choose the Cauchy slice ΣΣ\Sigma such that it intersects the boundary of the entanglement wedge at the QES. Let us also assume that ΣΣ\Sigma intersects the future causal horizon of DRsubscript𝐷𝑅D_{R} at H+(D)superscript𝐻𝐷H^{+}(D). The statement that the QES lies deeper than the causal surface can be proved by contradiction. Assume that χRIntH+(DR)subscript𝜒𝑅Intsuperscript𝐻subscript𝐷𝑅\chi_{R}\cap{\rm Int}H^{+}(D_{R}) is non empty as shown in Fig. 6. We can continuously shrink the boundary domain of dependence DRsubscript𝐷𝑅D_{R} to a new region Dsuperscript𝐷D^{\prime} such that the new causal horizon intersects ΣΣ\Sigma at H+(D)superscript𝐻superscript𝐷H^{+}(D^{\prime}), which is contained entirely in Ext(χR)Extsubscript𝜒𝑅{\rm Ext}(\chi_{R}) (see Fig.6). Since we are shrinking the region continuously we can choose a Dsuperscript𝐷D^{\prime} such that its future causal horizon H+(D)superscript𝐻superscript𝐷H^{+}(D^{\prime}) intersects χRsubscript𝜒𝑅\chi_{R} at p𝑝p, is tangent to it at p𝑝p and is in Ext(χR)Extsubscript𝜒𝑅{\rm Ext}(\chi_{R}) everywhere else. We can now identify H+(D)superscript𝐻superscript𝐷H^{+}(D^{\prime}) as N𝑁N from Eq. (15) to obtain the following:

δSgen(EW(χR))δΣμkμδSgen(H+(D))δΣμkμ𝛿subscript𝑆𝑔𝑒𝑛𝐸𝑊subscript𝜒𝑅𝛿superscriptΣ𝜇superscript𝑘𝜇𝛿subscript𝑆𝑔𝑒𝑛superscript𝐻superscript𝐷𝛿superscriptΣ𝜇superscript𝑘𝜇\frac{\delta S_{gen}(EW(\chi_{R}))}{\delta\Sigma^{\mu}}k^{\mu}\geq\frac{\delta S_{gen}(H^{+}(D^{\prime}))}{\delta\Sigma^{\mu}}k^{\mu} (16)

where EW(χR)𝐸𝑊subscript𝜒𝑅EW(\chi_{R}) (the boundary of the entanglement wedge) is the surface generated by shooting null rays from χRsubscript𝜒𝑅\chi_{R} towards R𝑅R. Since the QES extremizes the generalized entropy, the left hand side of Eq. (16) is zero. Therefore we have

δSgen(H+(D))δΣμkμ0,𝛿subscript𝑆𝑔𝑒𝑛superscript𝐻superscript𝐷𝛿superscriptΣ𝜇superscript𝑘𝜇0\frac{\delta S_{gen}(H^{+}(D^{\prime}))}{\delta\Sigma^{\mu}}k^{\mu}\leq 0, (17)

with equality only if χRsubscript𝜒𝑅\chi_{R} lies on H+(D)superscript𝐻superscript𝐷H^{+}(D^{\prime}) in a neighbourhood of p. However the generalized second law states that

δSgen(H+(D))δΣμkμ0.𝛿subscript𝑆𝑔𝑒𝑛superscript𝐻superscript𝐷𝛿superscriptΣ𝜇superscript𝑘𝜇0\frac{\delta S_{gen}(H^{+}(D^{\prime}))}{\delta\Sigma^{\mu}}k^{\mu}\geq 0. (18)

The equality holds only in non-generic spacetimes. Therefore for generic spacetimes we have a contradiction. A similar argument by using the time reversed GSL establishes a contradiction for the past causal horizon. The proof can be extended to the non generic case, see EngelhardtWall for details. Therefore the QES is spacelike or null separated from the causal surface and is deeper in the bulk.

Refer to caption
Figure 6: ΣΣ\Sigma is a spatial slice that contains the QES χRsubscript𝜒𝑅\chi_{R}. The projections of H+(D)superscript𝐻𝐷H^{+}(D) (blue) and H+(D)superscript𝐻superscript𝐷H^{+}(D^{\prime}) (red) are shown. The dotted curve is the QES that intersects and is tangent to H+(D)superscript𝐻superscript𝐷H^{+}(D^{\prime}) at point p. The region to the right of the QES is Ext(χR)𝐸𝑥𝑡subscript𝜒𝑅Ext(\chi_{R}). Figure reproduced from EngelhardtWall

This statement leads to an important conclusion. The von Neumann entropy is invariant under unitary transformations. Suppose we perturb the boundary with some unitary operator localized to R𝑅R, this should leave the boundary von Neumann entropy unchanged. This boundary perturbation leads to sources for the bulk fields. However only the bulk fields within WRsubscript𝑊𝑅W_{R} can be affected due to the boundary unitary. Since χRsubscript𝜒𝑅\chi_{R} lies deeper in the bulk than CRsubscript𝐶𝑅C_{R} and is spacelike or null separated from it, the QES is unchanged due to the boundary unitary perturbation. The bulk entanglement entropy Sent-bulksubscript𝑆ent-bulkS_{\text{ent-bulk}} is also unaffected since the von Neumann entropy is unchanged under unitary transformations. Thus the generalized entropy is unaffected by such boundary unitary perturbations. The classical extremal surface XRsubscript𝑋𝑅X_{R} can lie inside the causal wedge in spacetimes that violate the classical null energy condition. Thus the FLM holographic entropy would be changed under boundary unitaries, whereas the QES proposal is consistent with entanglement wedge reconstruction and the invariance of the boundary von Neumann entropy under unitary transformations of the state.

2.3 Maximin vs extremal: strong sub-additivity and entanglement wedge nesting

In the previous subsection we have described how the QES proposal is consistent with the invariance of the boundary von Neumann entropy under unitary transformations. There are two other conditions that any proposal for holographic entanglement entropy should satisfy: (1) strong sub-additivity of the von Neumann entropy and (2) entanglement wedge nesting. The second condition states that if we consider a boundary subregion RRsuperscript𝑅𝑅R^{\prime}\subset R then the entanglement wedge of Rsuperscript𝑅R^{\prime} should lie within the entanglement wedge of R𝑅R. The subalgebra of operators localized to Rsuperscript𝑅R^{\prime} must be a subset of the subalgebra of operators localized to R𝑅R and subregion duality implies the same must be true for the dual operators localized to the corresponding entanglement wedges. Thus entanglement wedge nesting is a consequence of subregion duality. The strong sub-additivity condition states that PhysRevLett.30.434 ; Lieb:1973cp

S(AB)+S(BC)S(B)+S(ABC),𝑆𝐴𝐵𝑆𝐵𝐶𝑆𝐵𝑆𝐴𝐵𝐶S(A\cup B)+S(B\cup C)\geq S(B)+S(A\cup B\cup C), (19)

where A𝐴A, B𝐵B and C𝐶C are three boundary sub-regions. A simple geometric proof of the strong sub-additivity for the RT prescription was given in Headrick:2007km (see Fig. 7).

Refer to caption
Figure 7: The left most figure computes S(AB)+S(BC)𝑆𝐴𝐵𝑆𝐵𝐶S(A\cup B)+S(B\cup C) via the RT prescription. The figure in the middle is simply a re-colouring of the different RT surfaces and must have the same entropy as the left most figure. The red and blue curves in the middle figure are not the extremal surfaces for the sub-regions B𝐵B and ABC𝐴𝐵𝐶A\cup B\cup C respectively, and therefore have a larger area than S(ABC)+S(B)𝑆𝐴𝐵𝐶𝑆𝐵S(A\cup B\cup C)+S(B). Figure from Nishioka:2009un .

Strong sub-additivity and entanglement wedge nesting was shown for the covariant HRT prescription in Wall:2012uf using the maximin surfaces defined as follows. For the boundary subregion R𝑅R we consider all possible Cauchy surfaces ΣΣ\Sigma that contain R𝑅R and find the minimal area surface that is homologous to R𝑅R (XR(Σ)subscript𝑋𝑅ΣX_{R}(\Sigma)) on each of these slices. Then we maximize over all such surfaces XR(Σ)subscript𝑋𝑅ΣX_{R}(\Sigma) to obtain the maximin surface. This surface is more convenient for proofs since it corresponds to a minimal surface on some Cauchy slice just like the RT surface. The maximin surface was shown to be equivalent to the HRT surface if the null curvature condition (NCC) holds Wall:2012uf . The NCC states that Rμνkμkν0subscript𝑅𝜇𝜈superscript𝑘𝜇superscript𝑘𝜈0R_{\mu\nu}k^{\mu}k^{\nu}\geq 0 for any null vector kμsuperscript𝑘𝜇k^{\mu}. Wall Wall:2012uf proved that these maximin surfaces exist in spacetimes without horizons and on spacetimes with Kasner like singularities. Thus in spacetimes satisfying the NCC the existence of HRT surfaces is guaranteed. This existence proof was then extended to generic blackholes in AdS𝐴𝑑𝑆AdS with singularities that are not Kasner like Marolf:2019bgj . The quantum generalization of the maximin surface was defined in Akers:2019lzs as follows. For a boundary subregion R𝑅R we consider all possible Cauchy surfaces ΣΣ\Sigma containing R𝑅R and find the surface that is homologous to R𝑅R and minimizes Sgensubscript𝑆𝑔𝑒𝑛S_{gen}, then we maximize over all the Cauchy slices. In Akers:2019lzs it was proved that the quantum maximin surfaces exist, are identical to the QES and obey strong sub-additivity as well as entanglement wedge nesting.

The RT/HRT and maximin prescription has led to stronger inequalities on the von Neumann entropy that do not hold for non holographic systems Hayden:2011ag ; Bao:2015bfa ; Hubeny:2018ijt ; Hubeny:2018trv . These inequalities haven’t been shown to hold when we include quantum corrections via the QES prescription, however exploration in this direction was initiated in Akers:2021lms where it was shown that if the bulk entropies obey the monogamy of mutual information Hayden:2011ag then the dual boundary entropies also obey the same.

3 Bulk reconstruction

3.1 The entanglement wedge reconstruction hypothesis

The AdS/CFT dictionary relates observables in the large N strongly coupled QFT at the boundary to the observables in the semi-classical bulk spacetime. The Euclidean partition function (Z𝑍Z) of the boundary theory is related to the on-shell bulk gravitational action (I𝐼I) as follows Witten:1998qj ; Gubser:1998bc :

Z[ϕ0]=eI[ϕ0],𝑍delimited-[]subscriptitalic-ϕ0superscript𝑒𝐼delimited-[]subscriptitalic-ϕ0Z[\phi_{0}]=e^{-I[\phi_{0}]}, (20)

where ϕ0subscriptitalic-ϕ0\phi_{0} is the boundary value of a bulk field ϕitalic-ϕ\phi and is identified with the source of the dual boundary operator 𝒪𝒪\mathcal{O}. Thus the bulk gravitational action is the generating functional of all connected correlation functions in the boundary theory. This is called the Gubser-Klebanov-Polyakov-Witten (GKPW) prescription in literature. This prescription implies that the connected correlation functions of the field theory can then be obtained by functionally differentiating the on-shell dual bulk gravitational action I𝐼I with respect to the sources. Divergences in the on-shell bulk gravitational action I𝐼I arise due to the infinite volume of the AdS𝐴𝑑𝑆AdS spacetime near the boundary and these mimic the local ultraviolet divergences in the dual field theory. These divergences can be systematically removed by first regularizing with a radial cut-off r=ϵ𝑟italic-ϵr=\epsilon (the boundary is at r=0𝑟0r=0) and subtracting them with diffeomorphism-invariant local counterterms on the cut-off surface. This procedure is called holographic renormalization Henningson:1998gx ; BalKraus ; deHaro:2000vlm ; Skenderis1 (see Kanitscheider:2008kd for implementation in more general cases). The radial cutoff thus mimics an energy-scale cutoff in the dual field theory. We discuss more on this issue in section 4.4. The Lorentzian generalization of (20) has been discussed in Son:2002sd ; Skenderis:2008dg ; Herzog:2002pc ; Glorioso:2018mmw .

An equivalent and useful way to state the correspondence which generalizes readily to the Lorentzian signature is as follows. Corresponding to any state ρBsubscript𝜌𝐵\rho_{B} in the boundary theory there exists an asymptotically AdSd+1𝐴𝑑subscript𝑆𝑑1AdS_{d+1} solution B𝐵B in the dual gravity theory which satisfies appropriate smoothness conditions such as absence of naked singularities (unless explicitly mentioned we will assume that B𝐵B has no horizon). The generic on-shell asymptotic boundary behavior of a scalar field ϕitalic-ϕ\phi dual to an operator 𝒪𝒪\mathcal{O} with scaling dimension ΔΔ\Delta in such a geometry is:

ϕ(r,t,𝐱)B=rdΔϕ0(t,𝐱)(1+𝒪(r2))+rΔO(t,𝐱)ρB(1+𝒪(r2)).italic-ϕsubscript𝑟𝑡𝐱𝐵superscript𝑟𝑑Δsubscriptitalic-ϕ0𝑡𝐱1𝒪superscript𝑟2superscript𝑟Δsubscriptdelimited-⟨⟩𝑂𝑡𝐱subscript𝜌𝐵1𝒪superscript𝑟2\phi(r,t,\mathbf{x})_{B}=r^{d-\Delta}\phi_{0}(t,\mathbf{x})(1+\mathcal{O}(r^{2}))+r^{\Delta}\langle O(t,\mathbf{x})\rangle_{\rho_{B}}(1+\mathcal{O}(r^{2})). (21)

The source ϕ0subscriptitalic-ϕ0\phi_{0} which couples to 𝒪𝒪\mathcal{O} is identified with the leading term (non-normalizable mode) in the asymptotic expansion as mentioned before. The coefficient of the sub-leading term (normalizable mode) gets identified with the expectation value of 𝒪𝒪\mathcal{O} in the dual state ρBsubscript𝜌𝐵\rho_{B} as indicated above. The mass of the field m𝑚m is related to the scaling dimension ΔΔ\Delta via

Δ=d2+d24+m2l2Δ𝑑2superscript𝑑24superscript𝑚2superscript𝑙2\Delta=\frac{d}{2}+\sqrt{\frac{d^{2}}{4}+m^{2}l^{2}}

with l𝑙l the AdS radius. An extrapolate dictionary stated in Susskind:1998dq ; Banks:1998dd relates correlation functions of the boundary theory in the state ρBsubscript𝜌𝐵\rho_{B} to scattering S𝑆S-matrices of the semiclassical bulk fields in the geometry B𝐵B as follows:555The right hand side of (22) is more precisely the AdS analogue of the Lehmann-Symanzik-Zimmermann (LSZ) reduction 1955NCimS…1..205L of S𝑆S-matrices in flat space.

𝒪(x1)𝒪(xn)ρB=limr0rnΔϕ(r,x1)ϕ(r,xn)B.subscriptdelimited-⟨⟩𝒪subscript𝑥1𝒪subscript𝑥𝑛subscript𝜌𝐵subscript𝑟0superscript𝑟𝑛Δsubscriptdelimited-⟨⟩italic-ϕ𝑟subscript𝑥1italic-ϕ𝑟subscript𝑥𝑛𝐵\left<\mathcal{O}(x_{1})\ldots\mathcal{O}(x_{n})\right>_{\rho_{B}}=\lim_{r\to 0}r^{-n\Delta}\langle\phi(r,x_{1})\ldots\phi(r,x_{n})\rangle_{B}. (22)

For instance a four point function O(x1)O(x2)O(x3)O(x4)delimited-⟨⟩𝑂subscript𝑥1𝑂subscript𝑥2𝑂subscript𝑥3𝑂subscript𝑥4\left<O(x_{1})O(x_{2})O(x_{3})O(x_{4})\right> can be obtained from a 22222\rightarrow 2 bulk scattering experiment shown in Fig. 8. This extrapolate dictionary has been shown to be equivalent to the GKPW prescription (20) in Giddings:1999qu ; Harlow:2011ke .

Refer to caption
Figure 8: A scattering experiment in the bulk that is equivalent to a 4 point function in the boundary. Figure from harlow2018tasi

The extrapolate dictionary reproduces boundary observables as boundary limits of bulk observables. However, this is not the goal of bulk reconstruction which aims to do the opposite, namely describe (reconstruct) bulk observables in terms of boundary observables. A naive way to do this is to solve the bulk equations of motion with boundary conditions determined by the data of the boundary CFT (expectation values of the dual operators etc) and then use the extrapolate dictionary. For a free bulk scalar field this gives (see Banks:1998dd for details of the computation):

ϕ(X)=ddxK(X,x)𝒪(x),italic-ϕ𝑋superscript𝑑𝑑𝑥𝐾𝑋𝑥𝒪𝑥\phi(X)=\int d^{d}xK(X,x)\mathcal{O}(x), (23)

where the integration is over all boundary points (x𝑥x) that are spacelike separated from the bulk point X𝑋X and K𝐾K is referred to as the smearing function (it is the inverse of the bulk-to-boundary propagator). This expression is correct at the leading order in N𝑁N and the 1/N1𝑁1/N corrections can be obtained by perturbatively solving the bulk equations of motion including the bulk vertices.

This naive procedure suffers from a major issue. Eq (23) says that a bulk local operator ϕ(X)italic-ϕ𝑋\phi(X) depends on all CFT operators localized to a region spacelike to X𝑋X. This non-locality persists even when the bulk operator is pushed to the boundary. Therefore Eq (23) doesn’t smoothly reduce to the extrapolate dictionary. In order to recover the extrapolate dictionary the smearing function K(X,x)𝐾𝑋𝑥K(X,x) must become more and more local as X𝑋X is pushed to the boundary.

Hamilton, Kabat, Lifschytz, and Lowe (HKLL) addressed this issue in the context of the AdS Rindler wedge which is the bulk causal wedge WRsubscript𝑊𝑅W_{R} of a ball-shaped region R𝑅R at the boundary. Recall that the boundary limit of WRsubscript𝑊𝑅W_{R}, i.e. WRsubscript𝑊𝑅\partial W_{R} is DRsubscript𝐷𝑅D_{R}, the boundary domain of dependence of R𝑅R. HKLL showed that the reconstruction of the bulk operator in a Rindler wedge can be made manifestly consistent with the extrapolate dictionary if we work in Rindler coordinates (which covers WRsubscript𝑊𝑅W_{R}) instead of the global coordinates of AdS HKLL ; HKLL2 . This is referred to as AdS-Rindler reconstruction. For example, we can choose the CFT state to be the vacuum. The dual geometry is then pure AdS. The HKLL reconstruction procedure is explicitly

ϕ(X)=DRdd1xdτKRindler(X;x,τ)𝒪(x,τ),italic-ϕ𝑋subscriptsubscript𝐷𝑅superscriptd𝑑1𝑥differential-d𝜏superscript𝐾Rindler𝑋𝑥𝜏𝒪𝑥𝜏\phi(X)=\int_{D_{R}}{\rm d}^{d-1}x{\rm d}\tau\,\,K^{\text{Rindler}}(X;x,\tau)\mathcal{O}(x,\tau), (24)

where τ𝜏\tau is the Rindler time. Note that the integration is restricted to DRsubscript𝐷𝑅D_{R} the boundary domain of dependence of R𝑅R and 𝒪(x,τ)𝒪𝑥𝜏\mathcal{O}(x,\tau) is the Heisenberg picture operator evolved with the Rindler Hamiltonian which generates translation in τ𝜏\tau, i.e. boosts. The smearing function KRindlersuperscript𝐾RindlerK^{\text{Rindler}} is known explicitly in terms of the mode functions obtained from the semi-classical quantization of the bulk field in Rindler wedge (see harlow2018tasi ). This leads to the causal wedge reconstruction conjecture, which states that a bulk field within the causal wedge WRsubscript𝑊𝑅W_{R} of a boundary subregion R𝑅R can be reconstructed on the boundary domain of dependence DRsubscript𝐷𝑅D_{R}, i.e. ϕ(X)italic-ϕ𝑋\phi(X) can be represented using boundary operators within DRsubscript𝐷𝑅D_{R} provided XWR𝑋subscript𝑊𝑅X\in W_{R}. As we move X𝑋X closer to the boundary, a smaller DRsubscript𝐷𝑅D_{R} is required to reconstruct ϕitalic-ϕ\phi on the boundary. This is manifestly consistent with the extrapolate dictionary and therefore solves the issue that occurred in the global reconstruction. It is important to note that the explicit smearing function is known only for the ball-shaped boundary subregions in the vacuum state.

The entanglement wedge reconstruction conjecture states that bulk operators within the entanglement wedge of some boundary region R𝑅R can be reconstructed from operators on DRsubscript𝐷𝑅D_{R} at the boundary. This is called the entanglement wedge reconstruction conjecture Czech_2012 ; Wall:2012uf ; FLM ; Headrick_2014 ; Jafferis_2016 . This conjecture has now been proven using methods of operator algebra (quantum) error correction as will be discussed in section 4.2. This automatically implies the causal wedge reconstruction since the causal wedge is contained within the entanglement wedge as shown earlier. Note that in the case of the AdS-Rindler wedge reconstruction, the causal and entanglement wedges coincide.

However, the prescription given by Eq. (24) cannot be correct when we consider entanglement wedge reconstruction. Generically the entanglement wedge contains a region which is spacelike separated from the boundary domain of dependence DRsubscript𝐷𝑅D_{R}. This is referred to as the causal shadow Headrick_2014 (see Fig. 9).666In Headrick_2014 , it was shown that a causal shadow is generated for an interval slightly larger than half the boundary (a circle) in the asymptotically AdS3𝐴𝑑subscript𝑆3AdS_{3} metric ds2=1cos2ρ(f(ρ)dt2+dρ2f(ρ)+sin2ρdϕ2),f(ρ)=112sin2(2ρ).formulae-sequencedsuperscript𝑠21superscript2𝜌𝑓𝜌dsuperscript𝑡2dsuperscript𝜌2𝑓𝜌superscript2𝜌dsuperscriptitalic-ϕ2𝑓𝜌112superscript22𝜌{\rm d}s^{2}=\frac{1}{\cos^{2}\rho}\left(-f(\rho){\rm d}t^{2}+\frac{{\rm d}\rho^{2}}{f(\rho)}+\sin^{2}\rho{\rm d}\phi^{2}\right),\quad f(\rho)=1-\frac{1}{2}\sin^{2}(2\rho). The metric is supported by bulk matter satisfying null energy condition. The causal shadow is spacelike to DRsubscript𝐷𝑅D_{R}, therefore all bulk operators in the causal shadow would commute with operators in DRsubscript𝐷𝑅D_{R}. Thus if an equation analogous to (24) is correct for entanglement wedge reconstruction, then all bulk operators in the causal shadow must commute with each other, which leads to an inconsistency since they are not necessarily mutually spacelike separated. See Fig. 9 for an illustration.

Refer to caption
Figure 9: The solid black lines are the edge of the boundary domain of dependence DRsubscript𝐷𝑅D_{R}. Solid blue lines are the edge of the causal wedge and the solid red lines are the edge of the entanglement wedge. The black dots are two bulk operators in the causal shadow region (beyond the causal wedge and within the entanglement wedge). These two operators need not be spacelike separated and need not commute, however a bulk reconstruction formula analogous to Eq.(24) would imply that they commute with each other.

Jafferis, Lewkowycz, Maldacena and Suh (JLMS) Jafferis_2016 proposed that this inconsistency can be resolved if the bulk reconstruction equation takes the form:

ϕ(X)italic-ϕ𝑋\displaystyle\phi(X) =Rdd1x𝑑sK(X;x,s)𝒪s(x),absentsubscript𝑅superscript𝑑𝑑1𝑥differential-d𝑠𝐾𝑋𝑥𝑠subscript𝒪𝑠𝑥\displaystyle=\int_{R}d^{d-1}x\int dsK(X;x,s)\mathcal{O}_{s}(x),
𝒪s(x)subscript𝒪𝑠𝑥\displaystyle\mathcal{O}_{s}(x) =ρRis𝒪(x)ρRis=eiHρs𝒪(x)ρRiHρs,absentsuperscriptsubscript𝜌𝑅𝑖𝑠𝒪𝑥superscriptsubscript𝜌𝑅𝑖𝑠superscript𝑒𝑖subscript𝐻𝜌𝑠𝒪𝑥superscriptsubscript𝜌𝑅𝑖subscript𝐻𝜌𝑠\displaystyle=\rho_{R}^{-is}\mathcal{O}(x)\rho_{R}^{is}=e^{iH_{\rho}s}\mathcal{O}(x)\rho_{R}^{-iH_{\rho}s}, (25)

where ρRsubscript𝜌𝑅\rho_{R} is the reduced density matrix on the boundary subregion R𝑅R and Hρ=logρsubscript𝐻𝜌𝜌H_{\rho}=-\log\rho is the modular Hamiltonian and K𝐾K is an appropriate smearing function. The conjugation of the operator 𝒪𝒪\mathcal{O} by the density matrix is called modular flow Haag:1992hx and s𝑠s is the modular flow parameter. Since the modular Hamiltonian Hρsubscript𝐻𝜌H_{\rho} is typically non-local, the modular flowed operators 𝒪ssubscript𝒪𝑠\mathcal{O}_{s} are non-local also, and therefore they do not commute with ϕitalic-ϕ\phi. This resolves the inconsistency due to the causal shadow described before. The JLMS proposal reduces to the HKLL prescription (24) for ball-shaped regions in the vacuum as the modular Hamiltonian is then exactly the Rindler Hamiltonian which generates boosts Bisognano:1976za ; Hislop1981 ; Casini:2011kv implying that 𝒪s(x)=𝒪(τ,x)subscript𝒪𝑠𝑥𝒪𝜏𝑥\mathcal{O}_{s}(x)=\mathcal{O}(\tau,x). Generically, the modular Hamiltonian is local only for boundary regions with sufficient symmetry and for vacuum states.

The entanglement wedge reconstruction hypothesis states that an appropriate smearing function K𝐾K should always exist for (3.1). Progress towards an explicit construction of this smearing function will be discussed later in this section.

3.2 The equivalence of bulk and boundary relative entropies and its consequences

We proceed to first show the key result of Jafferis_2016 which establishes the equivalence of the boundary and bulk relative entropies as a consequence of the Engelhardt-Wall prescription EngelhardtWall for the holographic entanglement entropy in the semi-classical approximation. This will be the fundamental input in the proof of entanglement wedge reconstruction to be discussed in section 4.2.2 in the framework of operator algebra error correction. We will also study some of the striking consequences which follow from this relation.

The relative entropy between two states ρ𝜌\rho and σ𝜎\sigma is a measure of their distinguishability (divergence) and is defined as

S(ρ|σ)=Tr[ρ(logρlogσ)]𝑆conditional𝜌𝜎Trdelimited-[]𝜌𝜌𝜎S(\rho|\sigma)={\rm Tr}[\rho(\log\rho-\log\sigma)] (26)

Its classical analogue is the Kullback-Leibler divergence between two probability distributions. It could be helpful to see how the relative entropy arises as a measure of distinguishability. Consider a positive-operator-valued-measure (POVM) which is a set of positive semi-definite operators Aisubscript𝐴𝑖A_{i} such that iAi=1^subscript𝑖subscript𝐴𝑖^1\sum_{i}A_{i}=\hat{1}. We then define the classical probability distributions p𝑝p and q𝑞q obtained via pi=Tr(Aiσ)subscript𝑝𝑖Trsubscript𝐴𝑖𝜎p_{i}={\rm Tr}(A_{i}\sigma) and qi=Tr(Aiρ)subscript𝑞𝑖Trsubscript𝐴𝑖𝜌q_{i}={\rm Tr}(A_{i}\rho), and use the classical Kullback-Liebler divergence between p𝑝p and q𝑞q to define

S1:=S1(ρ|σ)=supAi(ipi(logpilogqi))assignsubscript𝑆1subscript𝑆1conditional𝜌𝜎subscriptsupsubscript𝐴𝑖subscript𝑖subscript𝑝𝑖subscript𝑝𝑖subscript𝑞𝑖S_{1}:=S_{1}(\rho|\sigma)={\rm sup}_{A_{i}}\left(\sum_{i}p_{i}(\log p_{i}-\log q_{i})\right)

with the supremum taken over all possible POVMs. S1subscript𝑆1S_{1} is thus a measure of distinguishabilty between the two states for a single measurement. We can similarly consider n𝑛n copies of both ρ𝜌\rho and σ𝜎\sigma along with all POVMs acting on these n𝑛n-copies, and define Sn:=Sn(ρ|σ)assignsubscript𝑆𝑛subscript𝑆𝑛conditional𝜌𝜎S_{n}:=S_{n}(\rho|\sigma). The result of Hiai and Petz is that HiaiPetz :

S(ρ|σ)=limnSn.𝑆conditional𝜌𝜎subscript𝑛subscript𝑆𝑛S(\rho|\sigma)=\lim_{n\rightarrow\infty}S_{n}.

We can paraphrase this as the statement that the probability that we can confuse between ρ𝜌\rho and σ𝜎\sigma after we perform a large number (n𝑛n) of measurements on ρ𝜌\rho decreases as exp(nS(ρ|σ))𝑛𝑆conditional𝜌𝜎\exp(-nS(\rho|\sigma)) as n𝑛n\rightarrow\infty. In quantum field theory, the relative entropy is a measure of how well we can distinguish two states based on the algebra of observables in a subregion R𝑅R. See RevModPhys.90.045003 for a detailed and illuminating discussion. The relative entropy is invariant under simultaneous unitary transformations of the two states, and therefore like the von-Neumann entropy, it is an observable that depends only on DRsubscript𝐷𝑅D_{R} and not the specific choice of R𝑅R.

The first crucial property of the relative entropy is that it is non-negative and vanishes if and only if the two states are identical nielsen . It follows also from the similar feature of Kullback-Leibler divergence as should be clear from the above discussion. Furthermore, the relative entropy is related to mutual information. Consider the union of two subregions A𝐴A and B𝐵B, a joint state ρABsubscript𝜌𝐴𝐵\rho_{A\cup B} and the uncorrelated state ρAρBtensor-productsubscript𝜌𝐴subscript𝜌𝐵\rho_{A}\otimes\rho_{B} with each density matrix obtained by tracing out the complement of the corresponding subregion. The mutual information between A𝐴A and B𝐵B subregions in the joint state is defined as

I(A,B)=S(A)+S(B)S(AB)𝐼𝐴𝐵𝑆𝐴𝑆𝐵𝑆𝐴𝐵I(A,B)=S(A)+S(B)-S(A\cup B) (27)

with S(A)𝑆𝐴S(A), S(B)𝑆𝐵S(B) and S(AB)𝑆𝐴𝐵S(A\cup B) referring to the von Neumann entropies of ρAsubscript𝜌𝐴\rho_{A}, ρBsubscript𝜌𝐵\rho_{B} and ρABsubscript𝜌𝐴𝐵\rho_{A\cup B} respectively. One can readily see that

I(A,B)=S(ρAB|ρAρB).𝐼𝐴𝐵𝑆conditionalsubscript𝜌𝐴𝐵tensor-productsubscript𝜌𝐴subscript𝜌𝐵I(A,B)=S(\rho_{A\cup B}|\rho_{A}\otimes\rho_{B}). (28)

The non-negativity of the relative entropy then implies that I(A,B)𝐼𝐴𝐵I(A,B) is positive and vanishes only when the two intervals are fully uncorrelated.

The second crucial property of the relative entropy is its monotonicity under completely positive777Note that a map 𝒩:()(𝒦):𝒩𝒦\mathcal{N}:\mathcal{B}(\mathcal{H})\rightarrow\mathcal{B}(\mathcal{K}) between bounded linear operators in two Hilbert spaces \mathcal{H} and 𝒦𝒦\mathcal{K} is said to be positive if it maps positive operators on \mathcal{H} to positive operators on 𝒦𝒦\mathcal{K}. The map 𝒩𝒩\mathcal{N} is said to be completely positive if any extension of the map is also a positive map. In other words, suppose the map 𝒩𝒩\mathcal{N} acts on a subsystem Asubscript𝐴\mathcal{H}_{A} of a composite system ABtensor-productsubscript𝐴subscript𝐵\mathcal{H}_{A}\otimes\mathcal{H}_{B}, then complete positivity ensures that (𝒩)(ρAB)0tensor-product𝒩subscript𝜌𝐴𝐵0(\mathcal{N}\otimes\mathcal{I})(\rho_{AB})\geq 0, for all (positive) ρAB0subscript𝜌𝐴𝐵0\rho_{AB}\geq 0. trace preserving (CPTP) maps. A density matrix maps to another density matrix under a CPTP map and will characterize an arbitrary noise channel in the context of quantum error correction. It has been shown that uhlmann1977relative

S(𝒩(ρ)|𝒩(σ))S(ρ|σ)𝑆conditional𝒩𝜌𝒩𝜎𝑆conditional𝜌𝜎S(\mathcal{N}(\rho)|\mathcal{N}(\sigma))\leq S(\rho|\sigma) (29)

for an arbitrary CPTP map 𝒩𝒩\mathcal{N}. Considering 𝒩𝒩\mathcal{N} to be the tracing out of a subregion C𝐶C, we can readily see that the monotonicity of the relative entropy implies the strong subadditivity property (19) of the entanglement entropy as follows. Under this trace operation we should have the inequality

S(ρABC|ρAρBC)S(ρAB|ρAρB),𝑆conditionalsubscript𝜌𝐴𝐵𝐶tensor-productsubscript𝜌𝐴subscript𝜌𝐵𝐶𝑆conditionalsubscript𝜌𝐴𝐵tensor-productsubscript𝜌𝐴subscript𝜌𝐵S(\rho_{A\cup B\cup C}|\rho_{A}\otimes\rho_{B\cup C})\leq S(\rho_{A\cup B}|\rho_{A}\otimes\rho_{B}), (30)

from the monotonicity property. We easily obtain (19) from the above inequality using (28) and (27). The strong-subadditivity of the entanglement entropy is saturated for quantum Markov chain states in which A𝐴A and C𝐶C are independently conditioned by B𝐵B as will be discussed in section 4.2.2. This will have implications for toy models of holography.

To proceed further, we rewrite the relative entropy in the following form

S(ρ|σ)𝑆conditional𝜌𝜎\displaystyle S(\rho|\sigma) =\displaystyle= Tr[ρlogρσlogσ]+Tr[(σρ)logσ]𝑇𝑟delimited-[]𝜌𝜌𝜎𝜎𝑇𝑟delimited-[]𝜎𝜌𝜎\displaystyle Tr[\rho\log\rho-\sigma\log\sigma]+Tr[(\sigma-\rho)\log\sigma]
=\displaystyle= ΔS+ΔHσ,Δ𝑆Δdelimited-⟨⟩subscript𝐻𝜎\displaystyle-\Delta S+\Delta\left<H_{\sigma}\right>,

where ΔSΔ𝑆\Delta S is the difference between the von Neumann entropies of the states ρ𝜌\rho and σ𝜎\sigma, and ΔHσΔdelimited-⟨⟩subscript𝐻𝜎\Delta\left<H_{\sigma}\right> denotes the difference between the expectation value of the modular Hamiltonian of Hσ=logσsubscript𝐻𝜎𝜎H_{\sigma}=-\log\sigma in these two states. Note these differences are exact and not infinitesimal. Since the relative entropy is non-negative and reaches its extremal vanishing value when the two states are identical, the first order change in the relative entropy must vanish for an infinitesimal difference between the two states i.e. when ρ=σ+δσ𝜌𝜎𝛿𝜎\rho=\sigma+\delta\sigma. This implies the first law of entanglement entropy Blanco:2013joa

δS=δHσ𝛿𝑆𝛿delimited-⟨⟩subscript𝐻𝜎\delta S=\delta\left<H_{\sigma}\right> (31)

for any infinitesimal variation of the state σ𝜎\sigma.

The Engelhardt-Wall prescription states that for any boundary subregion R𝑅R

S(R)bdy=𝒜(χR)4G+Sent-bulk𝑆subscript𝑅bdy𝒜subscript𝜒𝑅4𝐺subscript𝑆ent-bulkS(R)_{\text{bdy}}=\frac{\mathcal{A}(\chi_{R})}{4G}+S_{\text{ent-bulk}} (32)

where χRsubscript𝜒𝑅\chi_{R} is the quantum extremal surface which extremizes the generalized entropy with Sent-bulksubscript𝑆ent-bulkS_{\text{ent-bulk}} the entanglement entropy of the bulk matter within the corresponding entanglement wedge. The area term can be viewed as the expectation value of an operator in the bulk effective field theory (Tr[ρA^χR4G]𝑇𝑟delimited-[]𝜌subscript^𝐴subscript𝜒𝑅4𝐺Tr[\rho\frac{\hat{A}_{\chi_{R}}}{4G}]). Note that the von-Neumann entropy is simply the expectation value of the modular Hamiltonian: S(σ)=Tr[σHσ]=Hσσ𝑆𝜎𝑇𝑟delimited-[]𝜎subscript𝐻𝜎subscriptdelimited-⟨⟩subscript𝐻𝜎𝜎S(\sigma)=Tr[\sigma H_{\sigma}]=\left<H_{\sigma}\right>_{\sigma}. Therefore Eq (32) can be written as an equivalence between the bulk and boundary modular Hamiltonians as Jafferis:2014lza :

Hbdy=A^χR4G+Hbulksubscript𝐻bdysubscript^𝐴subscript𝜒𝑅4𝐺Planck-constant-over-2-pisubscript𝐻bulkH_{\text{bdy}}=\frac{\hat{A}_{\chi_{R}}}{4G\hbar}+H_{\text{bulk}} (33)

Note that the area operator A^χRsubscript^𝐴subscript𝜒𝑅\hat{A}_{\chi_{R}} will commute with both Hbulksubscript𝐻bulkH_{\text{bulk}} and Hbdysubscript𝐻bdyH_{\text{bdy}} since XRsubscript𝑋𝑅X_{R} is spacelike separated with all points in the corresponding bulk and boundary regions. Furthermore, the boundary relative entropy in the form (3.2) is

S(ρ|σ)bdy=ΔSbdy+ΔHσbdy.𝑆subscriptconditional𝜌𝜎bdyΔsubscript𝑆bdyΔdelimited-⟨⟩superscriptsubscript𝐻𝜎bdyS(\rho|\sigma)_{\text{bdy}}=-\Delta S_{\text{bdy}}+\Delta\left<H_{\sigma}^{\text{bdy}}\right>. (34)

From (32) we readily obtain that

ΔSbdy=Δ(𝒜(XR)4G)+ΔSbulkΔsubscript𝑆bdyΔ𝒜subscript𝑋𝑅4𝐺Planck-constant-over-2-piΔsubscript𝑆bulk\Delta S_{\text{bdy}}=\Delta\left(\frac{\mathcal{A}(X_{R})}{4G\hbar}\right)+\Delta S_{\text{bulk}} (35)

and similarly from (33) we obtain

ΔHσbdy=ΔA^χR4G+ΔHσbulk.Δdelimited-⟨⟩superscriptsubscript𝐻𝜎bdyΔdelimited-⟨⟩subscript^𝐴subscript𝜒𝑅4𝐺Planck-constant-over-2-piΔdelimited-⟨⟩superscriptsubscript𝐻𝜎bulk\Delta\left<H_{\sigma}^{\text{bdy}}\right>=\Delta\left<\frac{\hat{A}_{\chi_{R}}}{4G\hbar}\right>+\Delta\left<H_{\sigma}^{\text{bulk}}\right>. (36)

Subtracting Eq. (35) from Eq (36) and noticing that the area term cancels, we immediately obtain the desired result Jafferis_2016

S(ρ|σ)bdy=S(ρ|σ)bulk𝑆subscriptconditional𝜌𝜎bdy𝑆subscriptconditional𝜌𝜎bulkS(\rho|\sigma)_{\text{bdy}}=S(\rho|\sigma)_{\text{bulk}} (37)

stating the equivalence between the boundary and bulk relative entropies. It is interesting to note that a similar equivalence between boundary and bulk mutual information (between two subregions) can also be readily proved FLM .

Let us consider a bulk field ϕitalic-ϕ\phi within the entanglement wedge of some boundary region R𝑅R. Eq. (33) implies:

[Hbdy,ϕ]=[Hbulk,ϕ]subscript𝐻bdyitalic-ϕsubscript𝐻bulkitalic-ϕ[H_{\text{bdy}},\phi]=[H_{\text{bulk}},\phi] (38)

The area term in Hbdysubscript𝐻bdyH_{\text{bdy}}, which is localized on the extremal surface and is spacelike to the interior of the entanglement wedge, drops out. Therefore bulk causality implies that the bulk and boundary modular flows are identical.

The definition of the modular Hamiltonian contains some ambiguities. For example in lattice gauge theories observables are located on the links of the lattice. Therefore, when a surface splitting space into two parts cuts a link, it is not clear which side of the surface should include the observable on that link. These ambiguities are localized on the boundary of R𝑅R. Nevertheless, the relative modular Hamiltonian Hrel-bdy=HRbdyHR¯bdysubscript𝐻rel-bdysubscript𝐻𝑅bdysubscript𝐻¯𝑅bdyH_{\text{rel-bdy}}=H_{R-\text{bdy}}-H_{\bar{R}-\text{bdy}} is free of such ambiguities. Similarly, the bulk modular Hamiltonian has ambiguities localized on the extremal surface, but the bulk relative modular Hamilton is free of these ambiguities. The relative bulk and boundary modular Hamiltonians should be identical, i.e. Hrel-bdy=Hrel-bulksubscript𝐻rel-bdysubscript𝐻rel-bulkH_{\text{rel-bdy}}=H_{\text{rel-bulk}}, if the area term cancels out, i.e. if R𝑅R and its complement R¯¯𝑅\overline{R} share the same extremal surface (and hence the two corresponding entanglement wedges are complements of each other). This happens for a pure state (and horizonless bulk geometries). This ambiguity free modular Hamiltonian can be used to define the modular flow.

To see consequences of the equivalence of bulk and boundary relative entropies, we need the first law of entanglement entropy (31) which simply follows from the vanishing of the first order variation of the relative entropy as shown above. We will also need the result for the second order variation of the relative entropy. Let

ρ=σ(ϵ)=σ+ϵδ1σ+ϵ2δ2σ+𝒪(ϵ3)𝜌𝜎italic-ϵ𝜎italic-ϵsubscript𝛿1𝜎superscriptitalic-ϵ2subscript𝛿2𝜎𝒪superscriptitalic-ϵ3\rho=\sigma(\epsilon)=\sigma+\epsilon\delta_{1}\sigma+\epsilon^{2}\delta_{2}\sigma+\mathcal{O}(\epsilon^{3})

. We readily see that

δS(σ(ϵ)|σ)=ϵ212Tr[δ1σddϵlog(σ+ϵδ1σ)|ϵ=0]+𝒪(ϵ3)𝛿𝑆conditional𝜎italic-ϵ𝜎superscriptitalic-ϵ212Trdelimited-[]evaluated-atsubscript𝛿1𝜎dditalic-ϵ𝜎italic-ϵsubscript𝛿1𝜎italic-ϵ0𝒪superscriptitalic-ϵ3\delta S(\sigma(\epsilon)|\sigma)=\epsilon^{2}\frac{1}{2}{\rm Tr}\left[\delta_{1}\sigma\frac{{\rm d}}{{\rm d}\epsilon}\log(\sigma+\epsilon\delta_{1}\sigma)\Big{|}_{\epsilon=0}\right]+\mathcal{O}(\epsilon^{3}) (39)

where terms containing δ2σsubscript𝛿2𝜎\delta_{2}\sigma vanish for the same reason as in the case of the first order variation mentioned before. This implies

d2S(σ(ϵ)|σ)dϵ2|ϵ=0=Tr[δ1σddϵlog(σ+ϵδ1σ)|ϵ=0]:=δ1σ,δ1σσevaluated-atsuperscriptd2𝑆conditional𝜎italic-ϵ𝜎dsuperscriptitalic-ϵ2italic-ϵ0Trdelimited-[]evaluated-atsubscript𝛿1𝜎dditalic-ϵ𝜎italic-ϵsubscript𝛿1𝜎italic-ϵ0assignsubscriptsubscript𝛿1𝜎subscript𝛿1𝜎𝜎\frac{{\rm d}^{2}S(\sigma(\epsilon)|\sigma)}{{\rm d}\epsilon^{2}}\Big{|}_{\epsilon=0}={\rm Tr}\left[\delta_{1}\sigma\frac{{\rm d}}{{\rm d}\epsilon}\log(\sigma+\epsilon\delta_{1}\sigma)\Big{|}_{\epsilon=0}\right]:=\left<\delta_{1}\sigma,\delta_{1}\sigma\right>_{\sigma} (40)

The quantity δ1σ,δ1σσsubscriptsubscript𝛿1𝜎subscript𝛿1𝜎𝜎\left<\delta_{1}\sigma,\delta_{1}\sigma\right>_{\sigma} is called the quantum Fisher information which defines a Riemannian metric on the space of states. Quantum Fisher information is important in the study of quantum metrology and state estimation, where it bounds the amount of information that can be obtained about a state by generalized measurements Petz2011 . The positivity of relative entropy implies that the quantum Fisher information is positive.

Let us consider σ𝜎\sigma to be the reduced density operator on a ball shaped subregion B𝐵B of the vacuum state of a holographic CFT and ρ𝜌\rho to be the corresponding reduced density operator of a perturbed state close to the CFT vacuum. The vacuum is dual to pure AdS and the perturbed state is dual to a perturbation of pure AdS. It was shown in Lashkari_2014 ; Faulkner_2014 using results from Casini:2011kv ; Hollands_2012 that at the leading and subleading orders the variation of the bulk relative entropy can be written in the form (with the bulk metric g=g0+ϵδg+𝒪(ϵ2)𝑔subscript𝑔0italic-ϵ𝛿𝑔𝒪superscriptitalic-ϵ2g=g_{0}+\epsilon\delta g+\mathcal{O}(\epsilon^{2})):888The key to this result is the map of the bulk Rindler wedge dual to the domain of dependence DBsubscript𝐷𝐵D_{B} of the ball shaped region at the boundary B𝐵B to a hyperbolic black hole usingCasini:2011kv (the Casini-Huerta-Myers (CHM) map). This map is dual to the statement that the vacuum state in DBsubscript𝐷𝐵D_{B} can be mapped conformally to a thermal density matrix in hyperbolic space with radius of curvature RH=1/(2πT)subscript𝑅𝐻12𝜋𝑇R_{H}=1/(2\pi T) given by the temperature. The perturbations of the vacuum is dual to gravitational perturbations of the hyperbolic black hole with a bifurcate horizon which can be analyzed by the method of Wald and Hollands in Hollands_2012 . Note that the QES is mapped to the bifurcate horizon. Then the change in bulk relative entropy can be split using (3.2) with the change in von-Neumann entropy given by the change of the black hole entropy and the change in the modular Hamiltonian (which gets identified with the usual Hamiltonian after the CHM map) given by the change in the Arnowitt-Misner-Deser energy. Then the change in the bulk relative entropy can be reproduced essentially from the variation of the gravitational action according to Hollands_2012 which further develops results in Iyer:1994ys .

δS(ρ|σ)bulk=(g0,δg,ξg)2ΣξμδEμν(g)𝑑Σν𝛿𝑆subscriptconditional𝜌𝜎bulksubscript𝑔0𝛿𝑔subscript𝜉𝑔2subscriptΣsuperscript𝜉𝜇𝛿subscript𝐸𝜇𝜈𝑔differential-dsuperscriptΣ𝜈\delta S(\rho|\sigma)_{\text{bulk}}=\mathcal{E}(g_{0},\delta g,\mathcal{L}_{\xi}g)-2\int_{\Sigma}\xi^{\mu}\delta E_{\mu\nu}(g)d\Sigma^{\nu} (41)

where ΣΣ\Sigma is the Cauchy slice bounded by the boundary subregion and the bulk extremal surface, δg𝛿𝑔\delta g is the bulk metric perturbation and

ξg=μξν+νξνsubscript𝜉𝑔subscript𝜇subscript𝜉𝜈subscript𝜈subscript𝜉𝜈\mathcal{L}_{\xi}g=\nabla_{\mu}\xi_{\nu}+\nabla_{\nu}\xi_{\nu}

is the Lie derivative of the bulk metric g𝑔g in the direction of ξ𝜉\xi which is the timelike Killing vector associated to the bulk Rindler wedge (note that ξ𝜉\xi vanishes on the extremal surface). Eμνsubscript𝐸𝜇𝜈E_{\mu\nu} is proportional to the equations of motion. Furthermore, \mathcal{E} is a symplectic form on ΣΣ\Sigma given by:

(g0,δ1g,δ2g)=116πΣδ1hμνδ2pμνδ2hμνδ1pμνsubscript𝑔0subscript𝛿1𝑔subscript𝛿2𝑔116𝜋subscriptΣsubscript𝛿1subscript𝜇𝜈subscript𝛿2superscript𝑝𝜇𝜈subscript𝛿2subscript𝜇𝜈subscript𝛿1superscript𝑝𝜇𝜈\mathcal{E}(g_{0},\delta_{1}g,\delta_{2}g)=-\frac{1}{16\pi}\int_{\Sigma}\delta_{1}h_{\mu\nu}\delta_{2}p^{\mu\nu}-\delta_{2}h_{\mu\nu}\delta_{1}p^{\mu\nu} (42)

where hh is the induced metric on ΣΣ\Sigma and pμν=h(KμνhμνK)superscript𝑝𝜇𝜈superscript𝐾𝜇𝜈superscript𝜇𝜈𝐾p^{\mu\nu}=\sqrt{h}(K^{\mu\nu}-h^{\mu\nu}K) with Kμνsubscript𝐾𝜇𝜈K_{\mu\nu} denoting the extrinsic curvature of the Cauchy surface (and K=hμνKμν𝐾superscript𝜇𝜈subscript𝐾𝜇𝜈K=h^{\mu\nu}K_{\mu\nu}).

Since ξg0=0subscript𝜉subscript𝑔00\mathcal{L}_{\xi}g_{0}=0 as ξ𝜉\xi is a Killing vector associated to the unperturbed metric g0subscript𝑔0g_{0}, the canonical energy vanishes, i.e. =00\mathcal{E}=0 at leading order in ϵitalic-ϵ\epsilon. The first order variation of relative entropy must vanish due to the positivity of relative entropy. Since this should hold for any Cauchy surface ΣΣ\Sigma, we obtain:

δEμν(g)=𝒪(ϵ2).𝛿subscript𝐸𝜇𝜈𝑔𝒪superscriptitalic-ϵ2\delta E_{\mu\nu}(g)=\mathcal{O}(\epsilon^{2}). (43)

This simply implies that the perturbed metric should satisfy linearized Einstein’s equation expanded about the background for an arbitrary perturbation. Therefore positivity of relative entropy is equivalent to linearized Einstein equations.

The second order variation of the relative entropy is then

d2S(ρ|σ)bulkdϵ2|ϵ=0=(δg,ξδg)2Σξμ2Eμν(g)ϵ2|ϵ=0dΣν.evaluated-atsuperscript𝑑2𝑆subscriptconditional𝜌𝜎bulk𝑑superscriptitalic-ϵ2italic-ϵ0𝛿𝑔subscript𝜉𝛿𝑔evaluated-at2subscriptΣsuperscript𝜉𝜇superscript2subscript𝐸𝜇𝜈𝑔superscriptitalic-ϵ2italic-ϵ0𝑑superscriptΣ𝜈\left.\frac{d^{2}S(\rho|\sigma)_{\text{bulk}}}{d\epsilon^{2}}\right|_{\epsilon=0}=\mathcal{E}(\delta g,\mathcal{L}_{\xi}\delta g)-2\int_{\Sigma}\xi^{\mu}\left.\frac{\partial^{2}E_{\mu\nu}(g)}{\partial\epsilon^{2}}\right|_{\epsilon=0}d\Sigma^{\nu}. (44)

The vanishing of the linearized equations of motion finally implies that Lashkari_2016 :

d2S(ρ|σ)bulkdϵ2|ϵ=0=(g0,δg,ξδg).evaluated-atsuperscript𝑑2𝑆subscriptconditional𝜌𝜎bulk𝑑superscriptitalic-ϵ2italic-ϵ0subscript𝑔0𝛿𝑔subscript𝜉𝛿𝑔\left.\frac{d^{2}S(\rho|\sigma)_{\text{bulk}}}{d\epsilon^{2}}\right|_{\epsilon=0}=\mathcal{E}(g_{0},\delta g,\mathcal{L}_{\xi}\delta g). (45)

The right hand side is the canonical energy of the linearized perturbation Hollands_2012 . Using the equivalence of the bulk and boundary relative entropies and (40), we obtain that the left hand side is exactly the quantum Fisher information in the boundary. Therefore, the quantum Fisher information of a perturbation of the density matrix in the CFT is dual to the bulk canonical energy of the dual linearized gravitational perturbation Lashkari_2016 . The positivity of the Fisher information (which follows from that of the relative entropy) then must imply the positivity of the canonical energy as indeed is the case for perturbation about any stable vacuum.

3.3 Modular flow and bulk reconstruction

3.3.1 The JLMS smearing function

In this section we describe the explicit construction of the smearing function Faulkner_2017 in equation 3.1. For this it is useful to first consider the Fourier transform of the modular flowed operators:

Oω=𝑑seisωeiHσsOeiHσs,[Hσ,Oω]=ωOω,formulae-sequencesubscript𝑂𝜔superscriptsubscriptdifferential-d𝑠superscript𝑒𝑖𝑠𝜔superscript𝑒𝑖subscript𝐻𝜎𝑠𝑂superscript𝑒𝑖subscript𝐻𝜎𝑠subscript𝐻𝜎subscript𝑂𝜔𝜔subscript𝑂𝜔O_{\omega}=\int_{-\infty}^{\infty}dse^{-is\omega}e^{iH_{\sigma}s}Oe^{-iH_{\sigma}s},\quad[H_{\sigma},O_{\omega}]=\omega O_{\omega}, (46)

where Hσsubscript𝐻𝜎H_{\sigma} is the modular Hamiltonian for the state σ𝜎\sigma. We describe below an explicit expression for the smearing function from Faulkner_2017 that can be obtained by looking at the zero modular frequency mode. We can consider a bulk operator ϕitalic-ϕ\phi in the entanglement wedge and look at its zero mode ϕ0(X)subscriptitalic-ϕ0𝑋\phi_{0}(X). Since the bulk modular flow is the same as boundary modular flow as seen in Eq.(38), it follows that [ϕω,Hσ]=ωϕωsubscriptitalic-ϕ𝜔subscript𝐻𝜎𝜔subscriptitalic-ϕ𝜔[\phi_{\omega},H_{\sigma}]=\omega\phi_{\omega}. Therefore the zero mode ϕ0subscriptitalic-ϕ0\phi_{0} commutes with the modular Hamiltonian and this field must be localized on the extremal surface. It was then shown in Faulkner_2017 that the zero mode of the dual boundary operator is:

O0(x)=χR𝑑XχRϕ(XχR)O(x)ϕ(XχR),subscript𝑂0𝑥subscriptsubscript𝜒𝑅differential-dsubscript𝑋subscript𝜒𝑅delimited-⟨⟩italic-ϕsubscript𝑋subscript𝜒𝑅𝑂𝑥italic-ϕsubscript𝑋subscript𝜒𝑅O_{0}(x)=\int_{\chi_{R}}dX_{\chi_{R}}\left<\phi(X_{\chi_{R}})O(x)\right>\phi(X_{\chi_{R}}), (47)

where χRsubscript𝜒𝑅\chi_{R} is the extremal surface corresponding to a boundary subregion R𝑅R. This generalizes the results from Czech:2016xec ; deBoer:2016pqk ; CarneirodaCunha:2016zmi . Inverting this expression gives

ϕ(XχR)=R𝑑xK0(XχR,x)O0(x),italic-ϕsubscript𝑋subscript𝜒𝑅subscript𝑅differential-d𝑥subscript𝐾0subscript𝑋subscript𝜒𝑅𝑥subscript𝑂0𝑥\phi(X_{\chi_{R}})=\int_{R}dxK_{0}(X_{\chi_{R}},x)O_{0}(x), (48)

where K0subscript𝐾0K_{0} can be obtained by inverting the usual bulk to boundary correlator defined as follows:

ϕ(XχR),O(x)=𝑑yK0(XχR,y)O(x),O0(y).italic-ϕsubscript𝑋subscript𝜒𝑅𝑂𝑥differential-d𝑦subscript𝐾0subscript𝑋subscript𝜒𝑅𝑦𝑂𝑥subscript𝑂0𝑦\left<\phi(X_{\chi_{R}}),O(x)\right>=\int dyK_{0}(X_{\chi_{R}},y)\left<O(x),O_{0}(y)\right>. (49)

This shows how bulk operators on the QES can be reconstructed. If we foliate the entanglement wedge by extremal surfaces corresponding to smaller and smaller boundary subregions contained within R𝑅R we can reconstruct operators on the full entanglement wedge. This foliation is well defined due to entanglement wedge nesting which was reviewed in section 2.3. The inversion of the bulk to boundary propagator only needs to be done over the extremal surface making this a much simpler computation compared to the inversion of the full bulk to boundary propagator. (A similar technical computation in a different context was done in Banerjee:2016mhh .)

These modular zero modes have been used to define the modular Berry connection for the boundary CFT Czech:2017zfq , which has been related to the Riemann curvature in the bulk Czech:2019vih .

3.3.2 A note on modular Hamiltonians for excited states and bulk reconstruction

We have described the JLMS proposal and an explicit construction of the smearing function. The only missing ingredient for entanglement wedge reconstruction is the modular Hamiltonian, which can be explicitly obtained for sufficiently symmetric boundary subregions of the vacuum state Bisognano:1976za ; Hislop1981 ; Casini:2011kv ; Faulkner:2016mzt ; Casini:2017roe . However in generic situations the modular Hamiltonian is a non-local object that is very hard to compute (for examples see Klich:2015ina ; Lashkari:2015dia ; Klich:2017qmt ). Using the Casini, Huerta and Myers map Casini:2011kv Sarosi and Ugajin S_rosi_2018 have described an explicit CFT construction of the modular Hamiltonian for ball shaped subregions in slightly excited states close to the vacuum. Let the excited state be ρ=σ+δσ𝜌𝜎𝛿𝜎\rho=\sigma+\delta\sigma, then the modular Hamiltonian is S_rosi_2018 :

Hρ=Hσ+n=1(1)n𝑑s1𝑑sn𝒦n(s1sn)𝒫subscript𝐻𝜌subscript𝐻𝜎superscriptsubscript𝑛1superscript1𝑛superscriptsubscriptdifferential-dsubscript𝑠1differential-dsubscript𝑠𝑛subscript𝒦𝑛subscript𝑠1subscript𝑠𝑛𝒫\displaystyle H_{\rho}=H_{\sigma}+\sum_{n=1}^{\infty}(-1)^{n}\int_{-\infty}^{\infty}ds_{1}\ldots ds_{n}\mathcal{K}_{n}(s_{1}\ldots s_{n})\mathcal{P}
𝒫=i=1ne(isi2π+12)H0δσe(isi2π+12)H0𝒫superscriptsubscriptproduct𝑖1𝑛superscript𝑒𝑖subscript𝑠𝑖2𝜋12subscript𝐻0𝛿𝜎superscript𝑒𝑖subscript𝑠𝑖2𝜋12subscript𝐻0\displaystyle\mathcal{P}=\prod_{i=1}^{n}e^{-\left(i\frac{s_{i}}{2\pi}+\frac{1}{2}\right)H_{0}}\delta\sigma e^{\left(i\frac{s_{i}}{2\pi}+\frac{1}{2}\right)H_{0}} (50)

This construction of the modular Hamiltonian doesn’t assume that the CFT is holographic. Sarosi and Ugajin S_rosi_2018 show using the results of Faulkner_2014 (valid for the Rindler wedge) that their construction of the modular Hamiltonian relates the quantum Fisher information to the canonical energy of an emergent bulk for any CFT without assuming either the RT/HRT formula in the bulk or the large N𝑁N limit for the CFT. The expressions for the kernel 𝒦nsubscript𝒦𝑛\mathcal{K}_{n} in Eq.(3.3.2) can be found in S_rosi_2018 , here we reproduce the kernels for n=1,2𝑛12n=1,2:

𝒦1(s1)subscript𝒦1subscript𝑠1\displaystyle\mathcal{K}_{1}(s_{1}) =1(2coshs12)2absent1superscript2subscript𝑠122\displaystyle=\frac{1}{(2\cosh{\frac{s_{1}}{2}})^{2}} (51)
𝒦2(s1,s2)subscript𝒦2subscript𝑠1subscript𝑠2\displaystyle\mathcal{K}_{2}(s_{1},s_{2}) =116πicoshs12coshs22sinhs2s12absent116𝜋𝑖subscript𝑠12subscript𝑠22subscript𝑠2subscript𝑠12\displaystyle=\frac{1}{16\pi}\frac{i}{\cosh{\frac{s_{1}}{2}}\cosh{\frac{s_{2}}{2}}\sinh{\frac{s_{2}-s_{1}}{2}}} (52)

Note that the kernel K1subscript𝐾1K_{1} is the same as the kernel seen in the twirled Petz map Junge_2018 ; cotler2019_univR described in section 4.2. See Kabat:2020oic ; Kabat:2021akg for an explicit evaluation of Eq. (3.3.2).

3.4 Why bulk reconstruction is quantum error correction

We have already seen hints of a connection between entanglement wedge reconstruction and quantum error correction. For instance Eq.(33) and (38) have been argued to be equivalent to the conditions for quantum error correction dong_2016_2 . Equation 3.3.2 has a structure reminiscent of the twirled Petz map Junge_2018 ; cotler2019_univR which will be reviewed in section 4.2. Moreover entanglement wedge reconstruction leads to an interesting puzzle Almheiri_2015 . Consider three boundary subregions A𝐴A,B𝐵B and C𝐶C, as shown in Fig. 10. The gray regions are the entanglement wedges for the corresponding boundary subregions. The dot indicates a bulk field ϕ(X)italic-ϕ𝑋\phi(X). This bulk field lies outside the entanglement wedges of each of the three boundary regions. However the field lies in the entanglement wedge of AB𝐴𝐵A\cup B, BC𝐵𝐶B\cup C and AC𝐴𝐶A\cup C. Thus the bulk operator ϕitalic-ϕ\phi can be reconstructed on AB𝐴𝐵A\cup B and must therefore commute with all operators on C𝐶C due to causality in the boundary theory. Similarly we can argue that ϕitalic-ϕ\phi must commute with all operators on A𝐴A and B𝐵B by reconstructing it on BC𝐵𝐶B\cup C and AC𝐴𝐶A\cup C respectively. Therefore the reconstructed bulk operator must be proportional to the identity since it commutes with all CFT operators. This inconsistency can be avoided if the three representations of the bulk operator on the three subregions are not the same. Thus the same bulk operator is encoded as different operators on the boundary subregions. Such redundant encoding is essential to quantum error correction. This connection between bulk reconstruction and quantum error correction will be described in detail in section 4.2.

Refer to caption
Figure 10: Three boundary subregions with their corresponding entanglement wedges are shown. The dot indicates a bulk operator. Figure from Almheiri_2015

3.5 A first look at islands

Following the QES proposal by Engelhardt and Wall EngelhardtWall , the Page curve Page_1993 ; Page_2013 for an evaporating black hole in AdS2𝐴𝑑subscript𝑆2AdS_{2} was computed in AEMM ; PeningtonQES . Similar models for black hole evaporation were studied in Mertens_2019 ; Rozali_2020 . The results from these models show that the semi-classical geometry can see features of unitarity (Page curve) in black hole evaporation. The information paradox is of course not resolved since it is still not clear how the information about the black hole interior is encoded in the radiation and how it can be decoded. Nevertheless this was an important step towards the resolution of the paradox.

The setup for these computations is Jackiw-Teitelboim JACKIW1985343 ; TEITELBOIM198341 ; almheiri2015models (JT) gravity with bulk matter described by a 1+1111+1 dimensional CFT. This is dual to a quantum dot. At some time t=0𝑡0t=0 the bulk boundary conditions are changed by coupling the dual quantum dot to a wire described by the same 1+1111+1 dimensional CFT as in the bulk but on a flat background without dynamical gravity (see Fig. 11). Therefore quanta can now flow across the boundary and the black hole in AdS2𝐴𝑑subscript𝑆2AdS_{2} starts evaporating. The QES can be explicitly computed in this setup since the bulk entanglement entropy is that of a two dimensional CFT which can be obtained using the methods of Cardy and Calabrese Calabrese:2009qy . It was shown in AEMM ; PeningtonQES that the QES has a phase transition which leads to the turning around of the Page curve of the black hole. At early times, the QES remains close to the bifurcation point of the original black hole horizon (before coupling to the bath) and starts moving outward towards the boundary. The von Neumann entropy of the dual quantum dot increases due to the emitted Hawking quanta. After the Page time, when the black hole and the radiation have same number of degrees of freedom, (a more precise definition of the Page time is in Section 5.1), a different extremal surface has minimal generalized entropy. This QES is located just inside the event horizon of the black hole and has a decreasing entropy, and thus implying that the the entanglement entropy of the quantum dot decreases. This is the desired feature of the Page curve if the full system has an unitary evolution (to be discussed later).

Although a Page curve was seen for the entropy of the quantum dot (dual to the evaporating black hole), the entropy of the Hawking radiation computed semi-classically in AEMM was shown to grow monotonically as seen in Hawking’s original computation. In section 5.1, we will analyze these models from the point of view of the full gravitational path integral which computes the Rènyi entropies of the Hawking radiation and explain how the naive semi-classical computation should be refined by including appropriate saddles which automatically reproduce the location of the QES and give results consistent with unitarity. Such computations however simplify remarkably in a so-called doubly holographic setup when the bulk matter comprising the Hawking quanta is itself holographic. So, we briefly discuss this below.

In the setup of AlmheiriQES the full two-dimensional quantum dot plus wire (bath) system has a three dimensional holographic dual. This can be though of as a locally AdS3𝐴𝑑subscript𝑆3AdS_{3} geometry with a dynamical boundary where the JT gravity theory is located (see Fig. 11). This is essentially the same as the setup in Randall_1999 ; Karch_2001 where the dynamical boundary was called the Planck brane.

Refer to caption
Figure 11: The doubly holographic setup from AlmheiriQES . A quantum dot (shown in black) described holographically by JT gravity with holographic bulk matter is brought in contact with the same bath holographic CFT without gravity. The boundary condition is such that the black hole in JT gravity can evaporate. The holographic dual of the full setup is locally AdS3𝐴𝑑subscript𝑆3AdS_{3} spacetime with a codimension one Planck brane where the JT theory lives.

The computation of the generalized entropy in this setup is simple since the bulk entanglement entropy can be computed via the original RT/HRT prescription in the dual three dimensional gravity theory to obtain:

Sgen(x)=ϕ(x)4G(2)+A(3)(Xx)4G(3),subscript𝑆𝑔𝑒𝑛𝑥italic-ϕ𝑥4superscript𝐺2Planck-constant-over-2-pisuperscript𝐴3subscript𝑋𝑥4superscript𝐺3Planck-constant-over-2-piS_{gen}(x)=\frac{\phi(x)}{4G^{(2)}\hbar}+\frac{A^{(3)}(X_{x})}{4G^{(3)}\hbar}, (53)

where x𝑥x is the location of the QES on the Planck brane and Xxsubscript𝑋𝑥X_{x} is the classical extremal surface in three dimensions that is anchored to x𝑥x at one end and the boundary of the semi-infinite interval in the bath system at the other (see Fig. 12). The first area term in this two dimensional case is simply the value of the dilaton (ϕitalic-ϕ\phi) from the JT theory at the location of the QES. The entanglement wedge for the black hole in this setup is shown in Fig. 12.

Refer to caption
Figure 12: The red segment in the figure on the left indicates where the bulk entanglement entropy must be computed from the point of view of the 2D gravity theory. The red region on the right is the late time entanglement wedge of the black hole in the 3D theory. The entanglement wedge of the bath is the grey region in the right figure, which includes an island behind the horizon.

If we try to compute the entanglement entropy of the bath CFT by usual methods, that is by tracing out everything to the left of the black dot in Fig. 12 (left), we would end up with an entropy that increases forever in time. This was seen in AEMM . (This naive computation is the coarse-grained entropy as will be defined later.) However in the doubly holographic setup the full CFT (bulk + bath) is dual to a three dimensional geometry. We should therefore use the RT/HRT prescription to compute the entanglement entropy of the bath. This results in an entanglement wedge for the bath that is exactly the complement of the entanglement wedge for the quantum dot (black hole) and hence they have the same entanglement entropies and Page curves as should be the case if the full system is a pure state and evolves unitarily. (The QES has exactly the same phase transition in the doubly holographic setup at Page time as in the computation in AEMM which is for a general CFT.) If we look from the two dimensional perspective after the Page time (Fig. 12 (left)) we see two disconnected pieces in the entanglement wedge of the bath (black curves in Fig. 12) forming islands. However if we look at this from the three dimensional perspective, the two islands are connected (Fig. 12 (right)). This is a realization of ER=EPR𝐸𝑅𝐸𝑃𝑅ER=EPR Maldacena_2013 paradigm in which the bath should be connected to the interior via a bridge in an extra dimension (wormhole) as a result of large amount of entanglement between them generated by the accumulation of semiclassical Hawking EPR (maximally entangled) pairs.

Based on the above computation the authors of AlmheiriQES proposed a new rule for computing entanglement entropies in setups that involve reference systems coupled to gravitational systems. This new rule for S(R)𝑆𝑅S(R) the entanglement entropy of a subregion R𝑅R in the bath (without gravity) coupled to a gravitating system is as follows:

S(R)=minIExt𝐼[Sent-bulk(RI)+A(I)4G].𝑆𝑅subscript𝐼𝐼Extdelimited-[]subscriptSent-bulkRIAI4GS(R)=\min_{I}\underset{I}{\rm Ext}\left[S_{\text{ent-bulk}}(R\cup I)+\frac{A(\partial I)}{4G}\right]. (54)

According to this rule, we should first extremize over the islands I𝐼I in the gravitating system and then minimize over the extrema. Sent-bulksubscript𝑆ent-bulkS_{\text{ent-bulk}} is the entanglement obtained from the semiclassical description of the system (in the doubly hologrpahic case it is given by the area of the RT surface). When R𝑅R is the entire bath (right of the black dot in Fig. 12), the first term in Eq.(54) equals the entropy of the black hole (B𝐵B) since the full state on IBR𝐼𝐵𝑅I\cup B\cup R is pure. The second term is simply the area of the shared QES. Therefore this new island rule gives the same entropy for the bath as that of the black hole leading to the same Page curve for both subsystems.

After Page time the position of the QES in all such setups geometrically realizes the Hayden-Preskill time for information mirroring Hayden_2007 (more discussion in Section 5.3.1) AEMM . More precisely, if some information is thrown into the blackhole post Page time then after the Hayden Preskill time the information crosses the QES. Therefore, the information escapes the entanglement wedge of the boundary and enters the island i.e. the entanglement wedge of the wire (Hawking radiation). It follows that the information thrown into the black hole can be recovered from the Hawking radiation after the Hayden-Preskill time.

These island computations in two dimensional gravity have been extended to higher dimensions in Almheiri_higher ; chen2020quantum ; chen2020quantum2 . It has also been shown that islands can extend outside of event horizons almheiri2019islands . The doubly holographic setups have been further analyzed in Chen_2020 where the Page transition has been studied after excising intervals in the bath CFT . Further studies have been done in Gautason:2020tmk ; Akers:2019nfi ; Hartman:2020swn ; Hollowood:2020cou ; Anegawa:2020ezn ; Hashimoto:2020cas ; Balasubramanian:2020hfs ; Alishahiha:2020qza ; Geng:2020qvw . Islands in the context of de-Sitter and cosmological spacetimes have been studied in Hartman:2020khs ; Chen:2020tes ; Krishnan:2020fer ; VanRaamsdonk:2020tlr ; Balasubramanian:2020xqf ; Sybesma:2020fxg ; Geng:2021wcq and also in Geng:2021iyq in the context of AdS/BCFT duality. The island rule (54) can be derived generally without invoking the doubly holographic setup as will be reviewed in section 5.1.

4 Holography and Quantum Error Correction

4.1 Preliminaries: Quantum Error Correction

Quantum error correction (QEC) is a mathematical framework that allows for partial or complete recovery of quantum information that is corrupted or lost by noise arising due to unwanted interactions of the quantum system with the environment gottesman_qec . Formally, such noise is modelled as a completely positive trace-preserving (CPTP) map 𝒩:()(𝒦):𝒩𝒦\mathcal{N}:\mathcal{B}(\mathcal{H})\rightarrow\mathcal{B}(\mathcal{K}) from the set of bounded linear operators on Hilbert space \mathcal{H} to the set of bounded linear operators on another Hilbert space 𝒦𝒦\mathcal{K} nielsen . Such a CPTP map on the system density operators can be described in terms of a set of Kraus operators {Ei}subscript𝐸𝑖\{E_{i}\}, as, 𝒩(ρ)=iEiρEi𝒩𝜌subscript𝑖subscript𝐸𝑖𝜌superscriptsubscript𝐸𝑖\mathcal{N}(\rho)=\sum_{i}E_{i}\rho E_{i}^{\dagger}. The operators Eisubscript𝐸𝑖E_{i} are often said to be the error operators associated with the noise map 𝒩𝒩\mathcal{N}.

Since the no-cloning theorem prevents perfect copying of an arbitrary quantum state wootters1982 , QEC aims to protect against the effects of noise by encoding the information into entangled states of a larger Hilbert space. Specifically, for an d𝑑d-dimensional quantum system with associated Hilbert space \mathcal{H}, an [[n,k]]delimited-[]𝑛𝑘[[n,k]] quantum code protects k𝑘k qudits by encoding them into a dksuperscript𝑑𝑘d^{k}-dimensional subspace 𝒞𝒞\mathcal{C} of the n𝑛n-qudit space nsuperscripttensor-productabsent𝑛\mathcal{H}^{\otimes n}. Throughout this discussion, we will assume that the noise acts identically and independently on each of the n𝑛n qudits that constitute the encoded space.

A QEC code 𝒞𝒞\mathcal{C} is said to correct perfectly for the noise 𝒩𝒩\mathcal{N}, iff there exists a CPTP map :(n)(𝒞):superscripttensor-productabsent𝑛𝒞\mathcal{R}:\mathcal{B}(\mathcal{H}^{\otimes n})\rightarrow\mathcal{B}(\mathcal{C}) – often called the recovery map – such that,

(𝒩)(ρ)=ρ,ρsuchthat𝒫ρ𝒫=ρ,formulae-sequence𝒩𝜌𝜌for-all𝜌suchthat𝒫𝜌𝒫𝜌(\mathcal{R}\circ\mathcal{N})(\rho)=\rho,\forall\rho\quad{\rm such}\;{\rm that}\quad\mathcal{P}\rho\mathcal{P}=\rho, (55)

where 𝒫𝒫\mathcal{P} is the projection map onto the codespace 𝒞𝒞\mathcal{C}. Note that the noise acting on the encoded state is now a map on the n𝑛n-qudit space, leading to single-qudit as well as multi-qudit errors. A given quantum code can only correct for some subset of these errors on the n𝑛n-qudit space, indicated by a third parameter called the distance t𝑡t of the code. Thus, an [[n,k,t]]delimited-[]𝑛𝑘𝑡[[n,k,t]] quantum code can correct perfectly for the loss of any set of tn𝑡𝑛t\leq n qudits, or, equivalently it can correct for arbitrary errors on upto t12𝑡12\frac{t-1}{2} qudits. Algebraic and information-theoretic conditions for perfect QEC are known knill_laflamme97 . The algebra of the Pauli operators has lead to the rich framework of stabilizer codes and topological QEC. We refer to the comprehensive review by Terhal et al. qec_review2015 for further details and references.

On the other hand, a quantum code 𝒞𝒞\mathcal{C} is said to correct approximately for the noise map 𝒩𝒩\mathcal{N}, iff there exists a (CPTP) recovery map \mathcal{R} such that,

(𝒩)(ρ)ρ,ρsuchthat𝒫ρ𝒫=ρ,formulae-sequence𝒩𝜌𝜌for-all𝜌suchthat𝒫𝜌𝒫𝜌(\mathcal{R}\circ\mathcal{N})(\rho)\approx\rho,\forall\rho\quad{\rm such}\;{\rm that}\quad\mathcal{P}\rho\mathcal{P}=\rho, (56)

where 𝒫𝒫\mathcal{P} is the projection map onto the codespace 𝒞𝒞\mathcal{C}. The performance of a QEC protocol (𝒞,)𝒞(\mathcal{C},\mathcal{R}) described by the pair of codespace 𝒞𝒞\mathcal{C} and recovery \mathcal{R}, is quantified by the fidelity function F𝐹F which is a measure of how close two quantum states are. The fidelity between a pair of states ρ,σ()𝜌𝜎\rho,\sigma\in\mathcal{B}(\mathcal{H}) is defined as

F(ρ,σ)=Trρ1/2σρ1/2.𝐹𝜌𝜎Trsuperscript𝜌12𝜎superscript𝜌12F(\rho,\sigma)={\rm Tr}\sqrt{\rho^{1/2}\sigma\rho^{1/2}}. (57)

Moving beyond states, the framework of QEC can be easily extended to operator error correction kribs_oqec_2005 ; operatorqec_nielsen2007 . Indeed, the operator QEC (OQEC) framework is the most relevant one in the context of bulk reconstruction of observables in holography, and it will be useful to describe it in some detail here. OQEC generalises the susbspace structure of standard QEC to a subsystem structure as follows. Suppose the system Hilbert space \mathcal{H} has a decomposition of the form ABCdirect-sumtensor-productsubscript𝐴subscript𝐵subscript𝐶\mathcal{H}\equiv\mathcal{H}_{A}\otimes\mathcal{H}_{B}\oplus\mathcal{H}_{C}, for some choice of Hilbert spaces Asubscript𝐴\mathcal{H}_{A}, Bsubscript𝐵\mathcal{H}_{B} and Csubscript𝐶\mathcal{H}_{C}. Then, Asubscript𝐴\mathcal{H}_{A} is said to be an error-correcting subsystem for a noise map 𝒩𝒩\mathcal{N} acting on \mathcal{H}, if, for all ρ𝜌\rho with support on Asubscript𝐴\mathcal{H}_{A} and σ𝜎\sigma with support on Bsubscript𝐵\mathcal{H}_{B}, there exists a recovery map \mathcal{R} such that (𝒩)(ρσ)=ρσ𝒩tensor-product𝜌𝜎tensor-product𝜌superscript𝜎(\mathcal{R}\circ\mathcal{N})(\rho\otimes\sigma)=\rho\otimes\sigma^{\prime}, for some σ()superscript𝜎\sigma^{\prime}\in\mathcal{B}(\mathcal{H}) . Physically, this implies that information stored in subsystem Asubscript𝐴\mathcal{H}_{A} can be recovered from the action of noise 𝒩𝒩\mathcal{N} by the recovery map \mathcal{R}. Note that we recover the structure of standard QEC codes when the system Bsubscript𝐵\mathcal{H}_{B} is trivial (one-dimensional).

The subsystem QEC structure can then be used to recover for an algebra of observables, via the framework of operator algebra quantum error correction beny_oqec2007 . Here, the focus is on identifying subspaces or more generally subsystems of the system Hilbert space \mathcal{H}, such that we can reliably recover observables X𝑋X that have support on the chosen subspace or subsystem. More generally, we may consider the Csuperscript𝐶C^{*}-algebra 𝒜𝒜\mathscr{A} of observables on \mathcal{H}999Note that the operator algebra of observables is closed under addition, multiplication and Hermitian conjugation, thus forming a Csuperscript𝐶C^{*}-algebra.. We will first formally state the necessary and sufficient conditions for an algebra of operators to be correctable under a noise map 𝒩𝒩\mathcal{N}.

Theorem 4.1 (Operator Algebra QEC Condition).

The algebra of observables 𝒜𝒜\mathscr{A} on a codespace 𝒞𝒞\mathcal{C} with projector P𝑃P is correctable against noise 𝒩𝒩\mathcal{N} with operators {Ei}subscript𝐸𝑖\{E_{i}\} if and only if [EiP,X]=0subscript𝐸𝑖𝑃𝑋0[E_{i}P,X]=0 for all errors Eisubscript𝐸𝑖E_{i} associated with the noise 𝒩𝒩\mathcal{N} and all observables X𝒜𝑋𝒜X\in\mathscr{A}.

4.2 Holography as QEC

We now proceed to restate the concept of bulk reconstruction in AdS/CFT in the language of quantum error correction. This connection has already been touched upon in Sec. 3, especially in the context of relating operators in the bulk spacetime to operators of the boundary CFT (Sec. 3.4). Early works in this direction demonstrated – using certain toy models and specific quantum codes – that the local operators in the bulk can be interpreted as encoded operators on certain subspaces of the states of the CFT at the boundary Almheiri_2015 ; happy_2015 . The theory of quantum error correction is then invoked to show that these encoded operators are naturally protected against erasures on the boundary by virtue of their entanglement structure.

4.2.1 Operator error correction and Bulk Reconstruction

The first concrete application of QEC was to explain the somewhat counter-intuitive property that emerges in the context of AdS-Rindler bulk reconstruction, namely that the same bulk operator ϕ(𝐱)italic-ϕ𝐱\phi({\bf x}) can be reconstructed on the union of different pairs of boundary subregions as shown in Fig. 10. The framework of QEC allows for this to happen in a non-trial manner, via a proposal that the same bulk operator is encoded as different operators on the boundary subregions via a suitable erasure QEC code Almheiri_2015 . Note that the erasure noise map is a specific example of a quantum noise map, wherein the information is either left unaffected or erased with certain probability. The idea of bulk reconstruction using a quantum erasure code can be made concrete via two simple toy examples. The first example is provided the 333-qutrit code qutrit , defined by the span of the following three qutrit (333-dimensional) quantum states.

|0~ket~0\displaystyle|\tilde{0}\rangle =\displaystyle= 13(|000+|111+|222)13ket000ket111ket222\displaystyle\frac{1}{\sqrt{3}}(|000\rangle+|111\rangle+|222\rangle)
|1~ket~1\displaystyle|\tilde{1}\rangle =\displaystyle= 13(|012+|120+|201)13ket012ket120ket201\displaystyle\frac{1}{\sqrt{3}}(|012\rangle+|120\rangle+|201\rangle)
|2~ket~2\displaystyle|\tilde{2}\rangle =\displaystyle= 13(|021+|102+|210).13ket021ket102ket210\displaystyle\frac{1}{\sqrt{3}}(|021\rangle+|102\rangle+|210\rangle). (58)

The erasure-correcting property of this code ensures that any three-qutrit state within the codespace can be reconstructed even if one of the three qutrits is lost or erased. Given an operator O𝑂O that acts on the single-qutrit space, one can always find a 333-qutrit encoded operator O~~𝑂\tilde{O} which acts in the same way on the 333-qutrit code subspace. The unique feature of this specific code is that there exist encoded operators (also referred to as logical operators in the literature) that have support only on two of the three qutrits that constitute the encoded space. Furthermore, it is then possible to identify sets of encoded operators that have the same action on the codespace, but have nontrivial support on different pairs of qutrits. This indeed captures the essence of the bulk reconstruction via the entanglement wedges shown in Fig. 10 and demonstrates that such redundant encoding of operators is indeed possible.

More generally, we may consider the larger encoded space \mathcal{H} to be partitioned into subsystems Esubscript𝐸\mathcal{H}_{E} and E¯subscript¯𝐸\mathcal{H}_{\bar{E}}, where Esubscript𝐸\mathcal{H}_{E} denotes the subsystem whose states get erased due to the noise and E¯subscript¯𝐸\mathcal{H}_{\bar{E}} denotes the subsystem whose states are unaffected by the erasure noise. Codespaces are then subspaces of the form 𝒞EE¯𝒞tensor-productsubscript𝐸subscript¯𝐸\mathcal{C}\subset\mathcal{H}_{E}\otimes\mathcal{H}_{\bar{E}}, such that any state |ψ~𝒞ket~𝜓𝒞|\tilde{\psi}\rangle\in\mathcal{C} can be recovered after erasure of the Esubscript𝐸\mathcal{H}_{E} subsytem. In the context of operator error correction, this translates to the statement that corresponding to an operator O𝑂O that acts on the codespace, there exists an operator OE¯subscript𝑂¯𝐸O_{\bar{E}} acting only on subsytem E¯subscript¯𝐸\mathcal{H}_{\bar{E}} such that

OE¯|ψ~=O|ψ~,|ψ~𝒞.formulae-sequencesubscript𝑂¯𝐸ket~𝜓𝑂ket~𝜓for-allket~𝜓𝒞O_{\bar{E}}|\tilde{\psi}\rangle=O|\tilde{\psi}\rangle,\;\forall|\tilde{\psi}\rangle\in\mathcal{C}. (59)

It is now easy to identify the set of operators O𝑂O which can be corrected by a given codespace 𝒞𝒞\mathcal{C}, using the operator algebra QEC condition in Thm. 4.1. Formally, the operator error correction property in Eq. 59 holds if and only if O𝑂O commutes with any operator XEsubscript𝑋𝐸X_{E} which acts only on Esubscript𝐸\mathcal{H}_{E}. In other words, Eq. 59 holds for a given operator O𝑂O and codespace 𝒞𝒞\mathcal{C}, if and only if,

ψ~|[O,XE]|ψ~=0,|ψ~𝒞.formulae-sequencequantum-operator-product~𝜓𝑂subscript𝑋𝐸~𝜓0for-allket~𝜓𝒞\langle\tilde{\psi}|[O,X_{E}]|\tilde{\psi}\rangle=0,\;\forall|\tilde{\psi}\rangle\in\mathcal{C}. (60)

It is useful to note that the set of operators that satisfy Eq. 60 form a *-subalgebra of the operators on the codespace 𝒞𝒞\mathcal{C}.

Refer to caption
Figure 13: Bulk-boundary factorization on a time slice and the entanglement wedge. The boundary CFT is factorized as AA¯tensor-productsubscript𝐴subscript¯𝐴\mathcal{H}_{A}\otimes\mathcal{H}_{\bar{A}}. The shaded region denotes the entanglement wedge Asubscript𝐴\mathcal{E}_{A} corresponding to the boundary region A𝐴A. asubscript𝑎\mathcal{H}_{a} is the Hilbert space of the bulk excitations in Asubscript𝐴\mathcal{E}_{A}. χAsubscript𝜒𝐴\chi_{A} denotes the RT extremal surface. Figure taken from dong_2016_2 .

4.2.2 Proof of entanglement wedge reconstruction via operator QEC

We proceed to prove entanglement wedge reconstruction using the framework of operator algebra error correction described above. The crucial input would be the JLMS result of the equivalence of bulk and boundary relative entropies Jafferis_2016 discussed in section 3.2.

Formally, following the discussion in dong_2016_2 , we may consider a factorization of the boundary CFT \mathcal{H} into AA¯tensor-productsubscript𝐴subscript¯𝐴\mathcal{H}_{A}\otimes\mathcal{H}_{\bar{A}} as shown in Fig. 13. Let Asubscript𝐴\mathcal{E}_{A} and A¯subscript¯𝐴\mathcal{E}_{\bar{A}} denote the associated entanglement wedges and let asubscript𝑎\mathcal{H}_{a} and a¯subscript¯𝑎\mathcal{H}_{\bar{a}} denote the Hilbert space of bulk excitations in Asubscript𝐴\mathcal{E}_{A} and A¯subscript¯𝐴\mathcal{E}_{\bar{A}} respectively. The codespace is a suitably chosen subspace of the CFT, with a natural factorization of the form 𝒞=aa¯𝒞tensor-productsubscript𝑎subscript¯𝑎\mathcal{C}=\mathcal{H}_{a}\otimes\mathcal{H}_{\bar{a}}. The JLMS proposal can then be stated as follows, for a pair of density operators ρA,σA(A)subscript𝜌𝐴subscript𝜎𝐴subscript𝐴\rho_{A},\sigma_{A}\in\mathcal{B}(\mathcal{H}_{A}) and a pair of density operators ρa,σasubscript𝜌𝑎subscript𝜎𝑎\rho_{a},\sigma_{a} acting on the space asubscript𝑎\mathcal{H}_{a} corresponding to the bulk subregion a𝑎a.

S(ρA|σA)=S(ρa|σa).𝑆conditionalsubscript𝜌𝐴subscript𝜎𝐴𝑆conditionalsubscript𝜌𝑎subscript𝜎𝑎S(\rho_{A}|\sigma_{A})=S(\rho_{a}|\sigma_{a}). (61)

Since the relative entropy between a pair of operators S(ρ|σ)𝑆conditional𝜌𝜎S(\rho|\sigma) vanishes if and only if ρ=σ𝜌𝜎\rho=\sigma, the above equality is identical to the statement that ρA=σAsubscript𝜌𝐴subscript𝜎𝐴\rho_{A}=\sigma_{A} would imply ρa=σasubscript𝜌𝑎subscript𝜎𝑎\rho_{a}=\sigma_{a} and vice-versa.

The relative entropy equivalence in Eq. (61) in conjunction with the operator algebra QEC condition in Eq. (60) leads to the following reconstruction theorem Dong_2016 .

Theorem 4.2 (Bulk reconstruction).

Consider any code subspace 𝒞𝒞\mathcal{C}\subset\mathcal{H} of a finite-dimensional Hilbert space101010The assumption that the Hilbert space \mathcal{H} is finite can be accomplished by imposing a UV cutoff in the CFT. \mathcal{H} with the factorization =AA¯tensor-productsubscript𝐴subscript¯𝐴\mathcal{H}=\mathcal{H}_{A}\otimes\mathcal{H}_{\bar{A}} and an operator O𝑂O that acts on 𝒞𝒞\mathcal{C}. Suppose there exists a factorization of the codespace into 𝒞=aa¯𝒞tensor-productsubscript𝑎subscript¯𝑎\mathcal{C}=\mathcal{H}_{a}\otimes\mathcal{H}_{\bar{a}} such that,

  • (i)

    the operator O𝑂O acts only on asubscript𝑎\mathcal{H}_{a}, and,

  • (ii)

    for any pair of pure states |Ψ,|Φ𝒞ketΨketΦ𝒞|\Psi\rangle,|\Phi\rangle\in\mathcal{C}, the reduced density operators ρA¯=TrA[|ΨΨ|]subscript𝜌¯𝐴subscriptTr𝐴delimited-[]ketΨbraΨ\rho_{\bar{A}}=\mathrm{Tr}_{A}[|\Psi\rangle\langle\Psi|], σA¯=TrA[|ΦΦ|]subscript𝜎¯𝐴subscriptTr𝐴delimited-[]ketΦbraΦ\sigma_{\bar{A}}=\mathrm{Tr}_{A}[|\Phi\rangle\langle\Phi|] ρa¯=Tra[|ΨΨ|]subscript𝜌¯𝑎subscriptTr𝑎delimited-[]ketΨbraΨ\rho_{\bar{a}}=\mathrm{Tr}_{a}[|\Psi\rangle\langle\Psi|] and σa¯=Tra[|ΦΦ|]subscript𝜎¯𝑎subscriptTr𝑎delimited-[]ketΦbraΦ\sigma_{\bar{a}}=\mathrm{Tr}_{a}[|\Phi\rangle\langle\Phi|] satisfy

    ρa¯=σa¯ρA¯=σA¯.subscript𝜌¯𝑎subscript𝜎¯𝑎subscript𝜌¯𝐴subscript𝜎¯𝐴\rho_{\bar{a}}=\sigma_{\bar{a}}\Rightarrow\rho_{\bar{A}}=\sigma_{\bar{A}}.

Then, there exists an operator OAsubscript𝑂𝐴O_{A} acting only on Asubscript𝐴\mathcal{H}_{A} such that its action on the codespace is the same as that of the operator O𝑂O. In other words,

OA|Ψ=O|Ψ,|Ψ𝒞.formulae-sequencesubscript𝑂𝐴ketΨ𝑂ketΨfor-allketΨ𝒞O_{A}|\Psi\rangle=O|\Psi\rangle,\,\forall\,|\Psi\rangle\in\mathcal{C}. (62)

We will now outline the proof strategy of Dong et al Dong_2016 here, for completeness. We first note that the operator algebra QEC condition in Thm. 4.1 implies that the bulk reconstruction in Eq. (62) follows if we can establish that the action of the operator O𝑂O commutes with the action of any XA¯subscript𝑋¯𝐴X_{\bar{A}} (with support only on subsystem A¯subscript¯𝐴\mathcal{H}_{\bar{A}}) on the codespace (see Eq. (60)). This can be shown easily for any Hermitian operator O𝑂O, and then extended to general operators by linearity. For any real number λ𝜆\lambda consider two states |Ψ,|Φ𝒞ketΨketΦ𝒞|\Psi\rangle,|\Phi\rangle\in\mathcal{C}, such that,

|Ψ=eiλO|Φ.ketΨsuperscript𝑒𝑖𝜆𝑂ketΦ|\Psi\rangle=e^{i\lambda O}|\Phi\rangle.

If assumption (i) of the theorem holds, the operator O𝑂O acts only on the subsystem asubscript𝑎\mathcal{H}_{a}. If we further assume that O𝑂O is Hermitian, the two states |ΨketΨ|\Psi\rangle and |ΦketΦ|\Phi\rangle are related by a unitary operator, so that,

ρa¯subscript𝜌¯𝑎\displaystyle\rho_{\bar{a}} =\displaystyle= Tra[|ΨΨ|]subscriptTr𝑎delimited-[]ketΨbraΨ\displaystyle\mathrm{Tr}_{a}[|\Psi\rangle\langle\Psi|] (63)
=\displaystyle= Tra[(eiλOIa¯)|ΦΦ|(eiλOIa¯)]subscriptTr𝑎delimited-[]tensor-productsuperscript𝑒𝑖𝜆𝑂subscript𝐼¯𝑎ketΦbraΦtensor-productsuperscript𝑒𝑖𝜆𝑂subscript𝐼¯𝑎\displaystyle\mathrm{Tr}_{a}[(e^{i\lambda O}\otimes I_{\bar{a}})|\Phi\rangle\langle\Phi|(e^{-i\lambda O}\otimes I_{\bar{a}})]
=\displaystyle= σa¯.subscript𝜎¯𝑎\displaystyle\sigma_{\bar{a}}.

Now if we use assumption (ii) – which is equivalent to the relative entropy condition in Eq. (61) – we have, ρA¯=σA¯subscript𝜌¯𝐴subscript𝜎¯𝐴\rho_{\bar{A}}=\sigma_{\bar{A}}. This in turn implies that the expectation value of any operator XA¯subscript𝑋¯𝐴X_{\bar{A}} acting only on subsystem A¯subscript¯𝐴\mathcal{H}_{\bar{A}} is the same for the two states |ΨketΨ|\Psi\rangle and |ΦketΦ|\Phi\rangle. Thus,

Ψ|XA¯|Ψquantum-operator-productΨsubscript𝑋¯𝐴Ψ\displaystyle\langle\Psi|X_{\bar{A}}|\Psi\rangle =\displaystyle= Φ|XA¯|Φquantum-operator-productΦsubscript𝑋¯𝐴Φ\displaystyle\langle\Phi|X_{\bar{A}}|\Phi\rangle
Φ|eiλOXA¯eiλO|Φabsentquantum-operator-productΦsuperscript𝑒𝑖𝜆𝑂subscript𝑋¯𝐴superscript𝑒𝑖𝜆𝑂Φ\displaystyle\Rightarrow\langle\Phi|e^{-i\lambda O}X_{\bar{A}}e^{i\lambda O}|\Phi\rangle =\displaystyle= Φ|XA¯|Φ.quantum-operator-productΦsubscript𝑋¯𝐴Φ\displaystyle\langle\Phi|X_{\bar{A}}|\Phi\rangle. (64)

Expanding Eq. (64) to order λ𝜆\lambda, we get the desired commutativity condition, namely,

Φ|[O,XA¯]|Φ=0,|Φ𝒞.formulae-sequencequantum-operator-productΦ𝑂subscript𝑋¯𝐴Φ0for-allketΦ𝒞\langle\Phi|[O,X_{\bar{A}}]|\Phi\rangle=0,\;\forall\;|\Phi\rangle\in\mathcal{C}.

Theorem 4.2 thus proves that bulk reconstruction is possible via the entanglement wedge prescription, provided there exists a codespace which can be factorized into the bulk subregion a𝑎a and its complement a¯¯𝑎\bar{a} in such a way that the corresponding density operators satisfy the relative entropy equality in Eq. (61). We refer to pastawski_preskill2017 for a more detailed exploration of the various connections between holography and operator algebra quantum error correction. Non-perturbative gravity corrections imply only an approximate recovery of the entanglement wedge is possible (with exponentially small errors and even in absence of horizons) and the generalization of the present discussion can be found in Gesteau:2021jzp 111111Approximate state dependent recovery is a necessary starting point in the presence of horizons as discussed later..

One final aspect of holographic QEC that we would like to highlight here is the important link between the erasure correcting properties of a holographic code and the strong subadditivity inequality. Consider a tripartite quantum state ρABC(ABC)subscript𝜌𝐴𝐵𝐶tensor-productsubscript𝐴subscript𝐵subscript𝐶\rho_{ABC}\in\mathcal{B}(\mathcal{H}_{A}\otimes\mathcal{H}_{B}\otimes\mathcal{H}_{C}). The strong subadditivity inequality (see (19)) states that,

S(ρABC)+S(ρB)S(ρAB)+S(ρBC).𝑆subscript𝜌𝐴𝐵𝐶𝑆subscript𝜌𝐵𝑆subscript𝜌𝐴𝐵𝑆subscript𝜌𝐵𝐶S(\rho_{ABC})+S(\rho_{B})\leq S(\rho_{AB})+S(\rho_{BC}). (65)

This inequality can be simply understood as the positivity of conditional mutual information I(A;C|B)𝐼𝐴conditional𝐶𝐵I(A;C|B) between subsystems A𝐴A and C𝐶C, given B𝐵B. It turns out that the quantum states that saturate this inequality have an interesting structure, as shown in hayden_SSA2004 . Equality of Eq. (65) implies that the Hilbert space Bsubscript𝐵\mathcal{H}_{B} can be decomposed as a direct sum of tensor products of the form B=iBi1Bi2subscript𝐵subscript𝑖tensor-productsubscriptsubscriptsuperscript𝐵1𝑖subscriptsubscriptsuperscript𝐵2𝑖\mathcal{H}_{B}=\sum_{i}\mathcal{H}_{B^{1}_{i}}\otimes\mathcal{H}_{B^{2}_{i}}, and that the tripartite state then has the block diagonal form,

ρABC=ipiρABi1ρBi2C.subscript𝜌𝐴𝐵𝐶subscript𝑖tensor-productsubscript𝑝𝑖subscript𝜌𝐴subscriptsuperscript𝐵1𝑖subscript𝜌subscriptsuperscript𝐵2𝑖𝐶\rho_{ABC}=\sum_{i}p_{i}\rho_{AB^{1}_{i}}\otimes\rho_{B^{2}_{i}C}. (66)

This structure essentially implies conditional independence of subsystems A𝐴A and C𝐶C, given subsytem B𝐵B. In other words the state ρABCsubscript𝜌𝐴𝐵𝐶\rho_{ABC} can be thought of as a quantum Markov chain, the quantum analogue of a classical Markov Chain ABC𝐴𝐵𝐶A\rightarrow B\rightarrow C. Furthermore, it was shown that if subsystem C𝐶C is erased (or traced out) the tripartite state ρABCsubscript𝜌𝐴𝐵𝐶\rho_{ABC} can be reconstructed from the marginal state ρABsubscript𝜌𝐴𝐵\rho_{AB} via the action of a recovery map :(B)(BC):subscript𝐵tensor-productsubscript𝐵subscript𝐶\mathcal{R}:\mathcal{B}(\mathcal{H}_{B})\rightarrow\mathcal{B}(\mathcal{H}_{B}\otimes\mathcal{H}_{C}) called the Petz map, described below in Sec. 4.3.1. Note that this recovery map is completely independent of subsystem A𝐴A.

Refer to caption
Figure 14: The 555-qubit code visualised as a tensor network happy_2015 . The white dots represent uncontracted indices on the boundary and the red dots represent the contracted tensor legs in the bulk.

In the holographic setting, this Markov chain structure can be applied to get a stronger statement of bulk recovery, as follows. Let the boundary region A𝐴A in Fig. 13 be made up of three disjoint regions A1A2A3subscript𝐴1subscript𝐴2subscript𝐴3A_{1}\cup A_{2}\cup A_{3}. Let A=A1A3superscript𝐴subscript𝐴1subscript𝐴3A^{\prime}=A_{1}\cup A_{3} denote the union of two such unconnected, disjoint regions. The full boundary can be viewed as the union of the three regions A¯AA2¯𝐴superscript𝐴subscript𝐴2\bar{A}\cup A^{\prime}\cup A_{2}. The question is, does there exists a recovery map that can correct for erasure of the region A2subscript𝐴2A_{2}, without invoking the complementary region A¯¯𝐴\bar{A}? The quantum Markov condition states that this is indeed possible if the boundary state is such that AAA3𝐴superscript𝐴subscript𝐴3A\rightarrow A^{\prime}\rightarrow A_{3} form a quantum Markov chain, which will saturate the inequality in Eq. (65). Then, there exists a recovery map :(A)(A):subscriptsuperscript𝐴subscript𝐴\mathcal{R}:\mathcal{B}(\mathcal{H}_{A^{\prime}})\rightarrow\mathcal{B}(\mathcal{H}_{A}) that recovers for the erasure of A2subscriptsubscript𝐴2\mathcal{H}_{A_{2}}, without involving the subsystem A¯subscript¯𝐴\mathcal{H}_{\bar{A}}. In such a situation, the erasure is said to locally correctable pastawski_preskill2017 . This observation has potential implications for decoding of the black hole interior, as discussed in Sec. 5.3.2.

For a discussion on how such tensor network models can reproduce correlation functions and entropy of a three-dimensional black hole geometry see Bhattacharyya:2016hbx .

4.2.3 Tensor network toy model for bulk reconstruction

While the operator algebra QEC framework provided an abstract proof of existence of bulk reconstruction in the AdS/CFT correspondence, in this section we review the concrete tensor network based toy model of holography from happy_2015 . A tensor network can be visualised as a graph with a set of vertices {Vx,x=1,2,,N}formulae-sequencesubscript𝑉𝑥𝑥12𝑁\{V_{x},\;x=1,2,\ldots,N\}, with a quantum state |Vxxketsubscript𝑉𝑥subscript𝑥|V_{x}\rangle\in\mathcal{H}_{x} associated with each vertex. An isometric tensor is any linear map T:xy:𝑇subscript𝑥subscript𝑦T:\mathcal{H}_{x}\rightarrow\mathcal{H}_{y} such that TT=Ixsuperscript𝑇𝑇subscript𝐼𝑥T^{\dagger}T=I_{x}, the identity operator on xsubscript𝑥\mathcal{H}_{x}. Specifically, T:|xijTij|yj:𝑇ketsubscript𝑥𝑖subscript𝑗subscript𝑇𝑖𝑗ketsubscript𝑦𝑗T:|x_{i}\rangle\rightarrow\sum_{j}T_{ij}|y_{j}\rangle, where {|xi}ketsubscript𝑥𝑖\{|x_{i}\rangle\} and {|yj}ketsubscript𝑦𝑗\{|y_{j}\rangle\} denote compete orthonormal bases for xsubscript𝑥\mathcal{H}_{x} and ysubscript𝑦\mathcal{H}_{y} respectively. The local Hilbert space xsubscript𝑥\mathcal{H}_{x} at each vertex could admit a tensor product decomposition of the form x=k=1nxk\mathcal{H}_{x}=\otimes_{k=1}^{n_{x}}\mathcal{H}_{k}. The number of tensor indices depends on the factorization structure of the input and output spaces. For instance, the action of the isometric map T:12y:𝑇tensor-productsubscript1subscript2subscript𝑦T:\mathcal{H}_{1}\otimes\mathcal{H}_{2}\rightarrow\mathcal{H}_{y} on the basis states can be represented as

|i1i2jTji2i1|yj.ketsubscript𝑖1subscript𝑖2subscript𝑗subscript𝑇𝑗subscript𝑖2subscript𝑖1ketsubscript𝑦𝑗|i_{1}i_{2}\rangle\rightarrow\sum_{j}T_{ji_{2}i_{1}}|y_{j}\rangle. (67)

An interesting property of such isometric tensors is that it is possible to reinterpret an input factor as an output factor, upto a rescaling. Thus, the tensor map in Eq. (67) can be recast as as a map T~:A1BA2:~𝑇subscriptsubscript𝐴1tensor-productsubscript𝐵subscriptsubscript𝐴2\tilde{T}:\mathcal{H}_{A_{1}}\rightarrow\mathcal{H}_{B}\otimes\mathcal{H}_{A_{2}} as,

|i1j,i2Tji2i1|yji2.ketsubscript𝑖1subscript𝑗subscript𝑖2subscript𝑇𝑗subscript𝑖2subscript𝑖1ketsubscript𝑦𝑗subscript𝑖2|i_{1}\rangle\rightarrow\sum_{j,i_{2}}T_{ji_{2}i_{1}}|y_{j}i_{2}\rangle.

In general, a tensor T𝑇T with n𝑛n indices, ranging over d𝑑d values represents a quantum state in an n𝑛n-fold tensor product space of d𝑑d-dimensional quantum systems.

|ψ=i1,i2,,inTi1i2in|i1i2in.ket𝜓subscriptsubscript𝑖1subscript𝑖2subscript𝑖𝑛subscript𝑇subscript𝑖1subscript𝑖2subscript𝑖𝑛ketsubscript𝑖1subscript𝑖2subscript𝑖𝑛|\psi\rangle=\sum_{i_{1},i_{2},\ldots,i_{n}}T_{i_{1}i_{2}\ldots i_{n}}|i_{1}i_{2}\ldots i_{n}\rangle.

A special class of tensors called perfect tensors lead to encoding isometries for quantum error correcting codes in the following sense. A perfect tensor with 2n2𝑛2n indices describes a pure state of 2n2𝑛2n quantum systems with the property that any subset of n𝑛n systems is maximally entangled with the complementary set of n𝑛n systems121212Such states are called Absolutely Maximally Entangled (AME) states.. Such a tensor can then be thought of as a linear map from 111 (d𝑑d-dimensional) system to 2n12𝑛12n-1 (d𝑑d-dimensional) systems, encoding a single quantum system is such a way that it is protected against the erasure of any n1𝑛1n-1 subsystems. In QEC terminology, a perfect tensor with 2n2𝑛2n indices corresponds to a [[2n1,1,n]]delimited-[]2𝑛11𝑛[[2n-1,1,n]] code, as exemplified by the well known [[5,1,3]]delimited-[]513[[5,1,3]] stabilizer code 5qubit_laflamme .

In the context of holography, a tensor network can be interpreted as a map from the bulk to the boundary in the following sense. If we imagine the quantum states associated with the vertices of a graph to correspond to perfect tensors, the edges of the graph are associated with contractions of the tensor “legs”. Contracted tensor indices can then be associated with bulk degrees of freedom and uncontracted vertices are associated with the boundary degrees of freedom, as depicted in Fig. 14.

The holographic pentagon code shown in Fig. 14 is the simplest example of a tensor network based holographic code, and provides a nice demonstration of exact bulk reconstruction. The code geometry comprises of a uniform tiling of a hyperbolic disc by pentagons, with four pentagons adjacent at each vertex. A perfect tensor with six legs is placed at the center of each pentagon, so that each tensor has one uncontracted index, indicated by the red dot in Fig. 14. All other interior legs are contracted. The uncontracted leg in the interior can be interpreted as an encoded input in the bulk to the tensor isometry and the uncontracted legs at the boundary (the white dots at the boundary in Fig. 14) interpreted as the physical outputs at the boundary. The entire system can thus be viewed as a tensor network that maps the input legs in the bulk to the output legs at the boundary.

We refer to happy_2015 for a detailed discussion of the error correction properties of the pentagon code. Suffice it to note that the erasure-correcting properties of the underlying 555-qubit code ensure that this toy model can achieve bulk reconstruction by accessing only a subregion of the boundary. A full description of the operator reconstruction involves some interesting techniques such as tensor pushing and leads to a so-called greedy entanglement wedge reconstruction. Finally, we note that the 333-qutrit code described in Eq. (58) is also a perfect tensor and can be viewed as a triangular holographic code.

Moving beyond holographic codes based on perfect tensors, toy models of holography have been proposed using random tensor networks hayden2016_rtn . Unlike perfect tensor codes which are based on fixed isometric tensors, these models involve projecting onto maximally entangled states via random projection operations. Such random tensor networks are known to successfully demonstrate several holographic properties harlow2017_RT and have been recently used to demonstrate approximate bulk reconstruction jia2020_petz_RTN via the universal recovery map described in Sec. 4.3.2 below. Infinite dimensional HaPPY codes have been discussed in Gesteau:2020hoz . Here, it has been shown that the infinite dimensional code fails to reproduce long range correlations at the boundary, which are necessary ingredients of a dual CFT; thus indicating limitations of this model for AdS/CFT.

4.3 Bulk-boundary reconstruction via quantum recovery maps

The original bulk reconstruction proposal presented in Sec. 3.4 relied on an exact equivalence of the bulk and boundary relative entropies, as stated in Eq. (37) (equivalently (61)). While such an exact equivalence of the bulk and boundary entropies maybe argued for in an asymptotic setting, in a finite regime, within the framework of bulk effective field theories, it is expected that these two entropies may only be approximately equal. (This will also be a crucial issue in the context of establishing black hole interior reconstruction as an universal subsystem recovery map. It will be discussed in section 5.2.) The question then arises as to whether the QEC-based bulk reconstruction argument can be extended to this case of approximate equivalence of the bulk and boundary relative entropies. It turns out that the right framework to consider in this case if that of approximate quantum error correction (AQEC), rather than the perfect QEC situation considered thus far. In this section, we review some recent works cotler2019_univR ; chen2020_petz ; jia2020_petz_RTN that formulate the bulk reconstruction argument using ideas and techniques from approximate QEC.

4.3.1 Approximate QEC and the Petz map

As defined in Eq. (56) above, approximate QEC extends the framework of QEC to allow for situations where the state is not perfectly recovered, but recovered with high enough fidelity after the action of the noise map. In the case of perfect QEC, code constructions rely on decomposing the noise operators in terms of Pauli operators and then using the structure of the Pauli algebra to identify good code subspaces. The recovery operation is a two-step process comprising error detection, followed by application of the appropriate Pauli operators to correct for the errors nielsen . The problem of finding good approximate QEC codes is in general much harder, since it requires a search over both code spaces 𝒞𝒞\mathcal{C} and recovery maps \mathcal{R}, such that (𝒩)(ρ)𝒩𝜌(\mathcal{R}\circ\mathcal{N})(\rho) is close in fidelity to the density operator ρ(𝒞)𝜌𝒞\rho\in\mathcal{B}(\mathcal{C}).

An important tool that emerged in this context is the idea of a near-optimal universal recovery map, namely, the Petz map ohya_petz , that can reverse the effect of the noise to a high degree of fidelity. Given a noise map 𝒩𝒩\mathcal{N} with associated error operators {Ei}subscript𝐸𝑖\{E_{i}\} and a code space 𝒞𝒞\mathcal{C}, the Petz map ρ,𝒩subscript𝜌𝒩\mathcal{R}_{\rho,\mathcal{N}} corresponding to any density operator ρ𝜌\rho with support on the codespace is defined in terms of its Kraus operators ρ,𝒩={(Rρ,𝒩)i}subscript𝜌𝒩subscriptsubscript𝑅𝜌𝒩𝑖\mathcal{R}_{\rho,\mathcal{N}}=\{(R_{\rho,\mathcal{N}})_{i}\}, as barnum_Knill2002 ,

(Rρ,𝒩)i=ρ1/2Ei(𝒩(ρ))1/2,subscriptsubscript𝑅𝜌𝒩𝑖superscript𝜌12superscriptsubscript𝐸𝑖superscript𝒩𝜌12(R_{\rho,\mathcal{N}})_{i}=\rho^{1/2}E_{i}^{\dagger}(\mathcal{N}(\rho))^{-1/2}, (68)

where the inverse is taken on the support of the positive operator 𝒩(ρ)=iEiρEi𝒩𝜌subscript𝑖subscript𝐸𝑖𝜌superscriptsubscript𝐸𝑖\mathcal{N}(\rho)=\sum_{i}E_{i}\rho E_{i}^{\dagger}. In other words, the action of the Petz map ρ,𝒩subscript𝜌𝒩\mathcal{R}_{\rho,\mathcal{N}} on an arbitrary density operator σ(𝒞)𝜎𝒞\sigma\in\mathcal{B}(\mathcal{C}) can be written as,

ρ,𝒩(σ)subscript𝜌𝒩𝜎\displaystyle\mathcal{R}_{\rho,\mathcal{N}}(\sigma) =\displaystyle= ρ1/2(iEi𝒩(ρ)1/2σ𝒩(ρ)1/2Ei)ρ1/2superscript𝜌12subscript𝑖superscriptsubscript𝐸𝑖𝒩superscript𝜌12𝜎𝒩superscript𝜌12subscript𝐸𝑖superscript𝜌12\displaystyle\rho^{1/2}\left(\sum_{i}E_{i}^{\dagger}\mathcal{N}(\rho)^{-1/2}\sigma\mathcal{N}(\rho)^{-1/2}E_{i}\right)\rho^{1/2} (69)
=\displaystyle= ρ1/2𝒩(𝒩(ρ)1/2σ𝒩(ρ)1/2)ρ1/2.superscript𝜌12superscript𝒩𝒩superscript𝜌12𝜎𝒩superscript𝜌12superscript𝜌12\displaystyle\rho^{1/2}\mathcal{N}^{\dagger}\left(\mathcal{N}(\rho)^{-1/2}\sigma\mathcal{N}(\rho)^{-1/2}\right)\rho^{1/2}.

Here, 𝒩superscript𝒩\mathcal{N}^{\dagger} denotes the dual to the map 𝒩𝒩\mathcal{N}, with Kraus operators {Ei}superscriptsubscript𝐸𝑖\{E_{i}^{\dagger}\}. Furthermore, it is easy to see that

ρ,𝒩𝒩(ρ)=ρ.subscript𝜌𝒩𝒩𝜌𝜌\mathcal{R}_{\rho,\mathcal{N}}\circ\mathcal{N}(\rho)=\rho.

Thus, ρ,𝒩subscript𝜌𝒩\mathcal{R}_{\rho,\mathcal{N}} is the map that reverses perfectly, the effect of the noise map 𝒩𝒩\mathcal{N} on the state ρ𝜌\rho.

In the context of approximate QEC, the Petz map defined in Eq. (68) was shown to be a universal, near-optimal recovery map, where optimality was characterized using the average entanglement fidelity barnum_Knill2002 . Subsequently, a variant of the Petz map – defined over a codespace 𝒞𝒞\mathcal{C}, rather than a specific state ρ𝜌\rho – has been shown to be a universal, near-optimal recovery map in terms of the worst-case fidelity approxQEC . In what follows we will survey some of the recent works cotler2019_univR ; chen2020_petz ; jia2020_petz_RTN that use the Petz map construction to demonstrate a robust, universal recovery map for bulk reconstruction.

4.3.2 Bulk reconstruction using the Petz map

The Petz map was originally conceived in the context of understanding the monotonicity of quantum relative entropy petz_monotonicity2003 ; hayden_SSA2004 . Under the action of a noise map 𝒩𝒩\mathcal{N}, the relative entropy between two states ρ,σ𝜌𝜎\rho,\sigma can never increase, that is,

S(ρ|σ)S(𝒩(ρ)|𝒩(σ)).𝑆conditional𝜌𝜎𝑆conditional𝒩𝜌𝒩𝜎S(\rho|\sigma)\geq S(\mathcal{N}(\rho)|\mathcal{N}(\sigma)). (70)

This is often referred to as Uhlmann’s theorem uhlmann1977relative in quantum information theory. Since the relative entropy S(ρ|σ)𝑆conditional𝜌𝜎S(\rho|\sigma) vanishes if and only if ρ=σ𝜌𝜎\rho=\sigma, it can be thought of as a measure of distance between quantum states. The difference between S(ρ|σ)𝑆conditional𝜌𝜎S(\rho|\sigma) and S(𝒩(ρ)|𝒩(σ))𝑆conditional𝒩𝜌𝒩𝜎S(\mathcal{N}(\rho)|\mathcal{N}(\sigma)) can thus be used to quantify the extent to which the noise map 𝒩𝒩\mathcal{N} corrupts the quantum system. It was subsequently realised that the monotonicity inequality also captures the extent of recoverability of the states ρ𝜌\rho and σ𝜎\sigma under noise 𝒩𝒩\mathcal{N}. Suppose there exists a recovery map \mathcal{R} that recovers the states ρ𝜌\rho and σ𝜎\sigma perfectly from the effects of the noise 𝒩𝒩\mathcal{N}, namely, (𝒩)(ρ)=ρ𝒩𝜌𝜌(\mathcal{R}\circ\mathcal{N})(\rho)=\rho and (𝒩)(σ)=σ𝒩𝜎𝜎(\mathcal{R}\circ\mathcal{N})(\sigma)=\sigma, then, the inequality in Eq. (70) is saturated. Interestingly, the converse is also true, and the specific form of the recovery map that saturates monotonicity is indeed given by the form of Petz map defined with respect to σ𝜎\sigma and 𝒩𝒩\mathcal{N} petz_monotonicity2003 :

σ,𝒩(.)=σ1/2𝒩(𝒩(σ)1/2(.)𝒩(σ)1/2)σ1/2.\mathcal{R}_{\sigma,\mathcal{N}}(.)=\sigma^{1/2}\mathcal{N}^{\dagger}\left(\mathcal{N}(\sigma)^{-1/2}(.)\mathcal{N}(\sigma)^{-1/2}\right)\sigma^{1/2}.

Note that this form is identical to the one in Eq. (69), except that this is the Petz map that recovers the state σ𝜎\sigma perfectly under the action of the noise 𝒩𝒩\mathcal{N}.

The fact that saturation of Eq. (70) is a necessary and sufficient condition for exact recoverability provides us with a nice information theoretic interpretation of the JLMS proposal for bulk construction. In essence, the JLMS condition in Eq. (61) can be thought of as a saturation of the monotonicity of the relative entropy between operators on the corresponding bulk and boundary subregions under the action of the erasure noise map on the complementary boundary region. This naturally begs the question of what happens when the monotonicity inequality is only approximately saturated. In the holographic context, this would imply that the bulk boundary relative relative entropies are only approximately equal, perhaps to leading order. Does there exist a universal recovery map in this case, which can achieve approximate bulk reconstruction despite having access to only a certain subregion of the boundary? Remarkably, it turns out that the answer to this question is in the affirmative, and there exists more than one construction of such a universal recovery map for approximate bulk reconstruction, based on the Petz map.

4.3.3 The twirled Petz map

One proposal for a universal recovery map comes from a time-averaged form of the Petz map, called the twirled Petz map. The twirled form of the Petz map is motivated by a recent result in quantum information theory, which relates the difference in quantum relative entropies before and after the action of a noise map to the fidelity between the ideal and noisy states. Formally, for any two states ρ,σ()𝜌𝜎\rho,\sigma\in\mathcal{B}(\mathcal{H}) on some Hilbert space \mathcal{H} and any noise map 𝒩𝒩\mathcal{N}, there exists a recovery map ~σ,𝒩subscript~𝜎𝒩\tilde{\mathcal{R}}_{\sigma,\mathcal{N}} such that Junge_2018 ,

S(ρ|σ)S(𝒩(ρ)|𝒩(σ))2logF(ρ,(~σ,𝒩𝒩)(ρ)).𝑆conditional𝜌𝜎𝑆conditional𝒩𝜌𝒩𝜎2𝐹𝜌subscript~𝜎𝒩𝒩𝜌S(\rho|\sigma)-S(\mathcal{N}(\rho)|\mathcal{N}(\sigma))\geq-2\log F(\rho,(\tilde{\mathcal{R}}_{\sigma,\mathcal{N}}\circ\mathcal{N})(\rho)). (71)

Here, F(.)F(.) is the fidelity function defined in Eq. (57) above. For a map 𝒩𝒩\mathcal{N} that saturates monotonicity of relative entropy, both LHS and RHS identically vanish, for in this case the recovery map ~σ,𝒩subscript~𝜎𝒩\tilde{\mathcal{R}}_{\sigma,\mathcal{N}} is simply the Petz map which satisfies σ,𝒩(ρ)=ρsubscript𝜎𝒩𝜌𝜌\mathcal{R}_{\sigma,\mathcal{N}}(\rho)=\rho. For a map that does not saturate the monotonicity inequality, the LHS of Eq. (71) quantifies the deviation from saturating monotonicity, whereas the RHS quantifies the extent to which the recovery map ~σ,𝒩subscript~𝜎𝒩\tilde{\mathcal{R}}_{\sigma,\mathcal{N}} recovers the state ρ𝜌\rho after the action of the noise. In essence, the above inequality states that the fidelity with which a recovery map can correct for the action of the noise 𝒩𝒩\mathcal{N} is bounded by how close the map 𝒩𝒩\mathcal{N} comes to saturating the monotonicity inequality.

Furthermore, an explicit form of the universal recovery map ~σ,𝒩subscript~𝜎𝒩\tilde{\mathcal{R}}_{\sigma,\mathcal{N}} for the case where monotonicity is not exactly saturated was also given in Junge_2018 , as,

~σ,𝒩(.)=dtβ(t)\displaystyle\tilde{\mathcal{R}}_{\sigma,\mathcal{N}}(.)=\int dt\beta(t)
σ1it2𝒩([𝒩(σ)]1+it2(.)[𝒩(σ)]1it2)σ1+it2,\displaystyle\sigma^{\frac{1-it}{2}}\mathcal{N}^{\dagger}\left([\mathcal{N}(\sigma)]^{\frac{-1+it}{2}}(.)[\mathcal{N}(\sigma)]^{\frac{-1-it}{2}}\right)\sigma^{\frac{1+it}{2}}, (72)

where β(t)=(π/2)(cosh(πt)+1)1𝛽𝑡𝜋2superscript𝜋𝑡11\beta(t)=(\pi/2)(\cosh(\pi t)+1)^{-1}. This twirled form of the Petz map then provides a natural choice for a universal recovery map in holography, where t𝑡t is interpreted as the boundary modular time. This was formalised in the work of cotler2019_univR , which we briefly review here.

The starting point for approximate bulk reconstruction will be the approximate form of the relative entropy equivalence from Jafferis_2016 .

S(ρA|σA)=S(ρa|σa)+𝒪(1N),𝑆conditionalsubscript𝜌𝐴subscript𝜎𝐴𝑆conditionalsubscript𝜌𝑎subscript𝜎𝑎𝒪1𝑁S(\rho_{A}|\sigma_{A})=S(\rho_{a}|\sigma_{a})+\mathcal{O}\left(\frac{1}{N}\right), (73)

where ρAsubscript𝜌𝐴\rho_{A}, σAsubscript𝜎𝐴\sigma_{A} are density operators on the boundary subregion A𝐴A in Fig. 13 and ρasubscript𝜌𝑎\rho_{a}, σasubscript𝜎𝑎\sigma_{a} are density operators on the corresponding entanglement wedge region a𝑎a in the bulk. In other words, the JLMS equality conditions holds upto leading order in the CFT gauge group rank N𝑁N. Interpreting Eq. (71) as approximate saturation of the monotonicity of the relative entropy requires the existence of a mapping ρaρAsubscript𝜌𝑎subscript𝜌𝐴\rho_{a}\rightarrow\rho_{A} of states from the entanglement wedge in the bulk to the boundary subregion A𝐴A. The AdS/CFT correspondence can be described via an isometry 𝒱:𝒞:𝒱𝒞\mathcal{V}:\mathcal{C}\rightarrow\mathcal{H} from the code subspace to the boundary CFT \mathcal{H}. The map 𝒩𝒩\mathcal{N} is simply the partial trace operation, which traces out the complementary boundary region A¯¯𝐴\bar{A}. Thus, for any ρa(a)subscript𝜌𝑎subscript𝑎\rho_{a}\in\mathcal{B}(\mathcal{H}_{a}), and any fixed, full-rank state σa¯(a¯)subscript𝜎¯𝑎subscript¯𝑎\sigma_{\bar{a}}\in\mathcal{B}(\mathcal{H}_{\bar{a}}), we have,

𝒩(ρa)=TrA¯[𝒱(ρaσa¯)𝒱].𝒩subscript𝜌𝑎subscriptTr¯𝐴delimited-[]𝒱tensor-productsubscript𝜌𝑎subscript𝜎¯𝑎superscript𝒱\mathcal{N}(\rho_{a})=\mathrm{Tr}_{\bar{A}}[\mathcal{V}(\rho_{a}\otimes\sigma_{\bar{a}})\mathcal{V}^{\dagger}]. (74)

Note that we only consider product density operators of the form ρaσa¯tensor-productsubscript𝜌𝑎subscript𝜎¯𝑎\rho_{a}\otimes\sigma_{\bar{a}} on the codespace, where σa¯subscript𝜎¯𝑎\sigma_{\bar{a}} is a fixed state on a¯subscript¯𝑎\mathcal{H}_{\bar{a}}, so that the partial trace operation truly becomes a mapping of states ρasubscript𝜌𝑎\rho_{a} on the entanglement wedge region (shaded region a𝑎a in Fig. 13) to states TrA¯[𝒱(ρaσa¯)𝒱]subscriptTr¯𝐴delimited-[]𝒱tensor-productsubscript𝜌𝑎subscript𝜎¯𝑎superscript𝒱\mathrm{Tr}_{\bar{A}}[\mathcal{V}(\rho_{a}\otimes\sigma_{\bar{a}})\mathcal{V}^{\dagger}] on the boundary subregion A𝐴A.

Now, consider the twirled Petz map corresponding to the map 𝒩𝒩\mathcal{N} defined in Eq. (74), ~σa,𝒩subscript~subscript𝜎𝑎𝒩\tilde{\mathcal{R}}_{\sigma_{a},\mathcal{N}}, defined using a full-rank state σa(a)subscript𝜎𝑎subscript𝑎\sigma_{a}\in\mathcal{B}(\mathcal{H}_{a}). This is a map of the form ~σa,𝒩:(A)(a):subscript~subscript𝜎𝑎𝒩subscript𝐴subscript𝑎\tilde{\mathcal{R}}_{\sigma_{a},\mathcal{N}}:\mathcal{B}(\mathcal{H}_{A})\rightarrow\mathcal{B}(\mathcal{H}_{a}). Finally, if we assume that the map 𝒩𝒩\mathcal{N} defined in Eq. (74) approximately saturates the monotonicity inequality, Eq. (71) implies that the twirled Petz map ~σa,𝒩subscript~subscript𝜎𝑎𝒩\tilde{\mathcal{R}}_{\sigma_{a},\mathcal{N}} recovers any ρa(𝒞)subscript𝜌𝑎𝒞\rho_{a}\in\mathcal{B}(\mathcal{C}) with a fidelity that is bounded by,

2logF(ρa,(~σa,𝒩𝒩)(ρa))2𝐹subscript𝜌𝑎subscript~subscript𝜎𝑎𝒩𝒩subscript𝜌𝑎\displaystyle-2\log F(\rho_{a},(\tilde{\mathcal{R}}_{\sigma_{a},\mathcal{N}}\circ\mathcal{N})(\rho_{a})) (75)
\displaystyle\leq S(ρa|σa)S(𝒩(ρa)|𝒩(σa)).𝑆conditionalsubscript𝜌𝑎subscript𝜎𝑎𝑆conditional𝒩subscript𝜌𝑎𝒩subscript𝜎𝑎\displaystyle S(\rho_{a}|\sigma_{a})-S(\mathcal{N}(\rho_{a})|\mathcal{N}(\sigma_{a})).

This argument was then extended in cotler2019_univR to show that the map ~σa,𝒩subscript~subscript𝜎𝑎𝒩\tilde{\mathcal{R}}_{\sigma_{a},\mathcal{N}} can recover for all states, extending the scope of this result beyond states that are factorised as ρ=ρaσa¯𝜌tensor-productsubscript𝜌𝑎subscript𝜎¯𝑎\rho=\rho_{a}\otimes\sigma_{\bar{a}}. Furthermore, since ~σa,𝒩subscript~subscript𝜎𝑎𝒩\tilde{\mathcal{R}}_{\sigma_{a},\mathcal{N}} recovers states on asubscript𝑎\mathcal{H}_{a} with high fidelity, it can be shown that the adjoint ~σa,𝒩subscriptsuperscript~subscript𝜎𝑎𝒩\tilde{\mathcal{R}}^{\dagger}_{\sigma_{a},\mathcal{N}} maps operators ϕasubscriptitalic-ϕ𝑎\phi_{a} with support on the entanglement wedge region asubscript𝑎\mathcal{H}_{a} to boundary operators 𝒪Asubscript𝒪𝐴\mathcal{O}_{A} which are close in expectation values. Finally, it was shown in cotler2019_univR that an explicit formula for operators 𝒪Asubscript𝒪𝐴\mathcal{O}_{A} can be obtained by a specific choice of the fixed states σasubscript𝜎𝑎\sigma_{a} and σa¯subscript𝜎¯𝑎\sigma_{\bar{a}}, namely, the maximally mixed states on asubscript𝑎\mathcal{H}_{a} and a¯subscript¯𝑎\mathcal{H}_{\bar{a}}.

We conclude this section by noting a few more recent results that demonstrate bulk reconstruction using variants of the Petz map. Moving away from the twirled Petz map which involves an averaging over the modular time, it was argued in chen2020_petz that the standard form of the Petz map in Eq. (68) suffices to achieve approximate bulk reconstruction. Their argument is based on the original result of Barnum and Knill barnum_Knill2002 , which shows that the Petz map is the near-optimal universal recovery map in terms of the average entanglement fidelity between the ideal and noisy states. We restate the main result of chen2020_petz here, for completeness. Let asubscript𝑎\mathcal{M}_{a} be a subalgebra on the code space 𝒞𝒞\mathcal{C} with dimension dcodesubscript𝑑coded_{\rm code}, and 𝒩𝒩\mathcal{N} be any noise map. Suppose there exists an optimal recovery map optsubscriptopt\mathcal{R}_{\rm opt} such that

(opt𝒩)(ρ)ρa1<δ,subscriptnormsubscriptopt𝒩𝜌subscript𝜌𝑎1𝛿\parallel(\mathcal{R}_{\rm opt}\circ\mathcal{N})(\rho)-\rho_{a}\parallel_{1}<\delta,

for all ρ(𝒞)𝜌𝒞\rho\in\mathcal{B}(\mathcal{C}) and its projection ρasubscript𝜌𝑎\rho_{a} onto asubscript𝑎\mathcal{M}_{a}. Then, the Petz map

τ,𝒩(.)1dcode𝒩[𝒩(τ)1/2(.)𝒩(τ)1/2],\mathcal{R}_{\tau,\mathcal{N}}(.)\equiv\frac{1}{d_{\rm code}}\mathcal{N}^{\dagger}\left[\mathcal{N}(\tau)^{-1/2}(.)\mathcal{N}(\tau)^{-1/2}\right],

defined using the maximally mixed state τ𝜏\tau on the codespace, satisfies,

(τ,𝒩)𝒩)(ρ)|aρa1dcode8δ,\parallel(\mathcal{R}_{\tau,\mathcal{N}})\circ\mathcal{N})(\rho)|_{a}-\rho_{a}\parallel_{1}\leq d_{\rm code}\sqrt{8\delta},

where (.)1\parallel(.)\parallel_{1} denotes the 111-norm or the trace-distance and (.)|a(.)|_{a} denotes the projection onto the subalgebra asubscript𝑎\mathcal{M}_{a}. In other words, so long as the error using the optimal recovery opsubscriptop\mathcal{R}_{\rm op} is non-perturbatively small (in trace-norm), the error after the Petz map recovery will also be non-perturbatively small upto a factor dcodesubscript𝑑coded_{\rm code}.

More recently, it has been shown that a Petz-like map can be used to demonstrate bulk reconstruction in random tensor network toy models jia2020_petz_RTN ; Penington:2019kki of holography. From a dynamical point of view, there appears to be an interesting connection between the structure of the twirled Petz map and the action of the modular Hamiltonian, as hinted in cotler2019_univR and also in sections 3.3.2 and 3.4. Exploring this connection further might lead to further insights on the problem of bulk reconstruction, and promises to be an exciting direction for future investigations.

4.4 A note on holography as a renormalization group flow

Holographic renormalization scheme which defines the dictionary between boundary and bulk observables already makes it manifest that the radial direction in the holographic bulk is dual to the energy scale/scale of resolution in the dual field theory. Nevertheless, it does not in itself tell us how local bulk operators such as the metric can be reconstructed as coarse-grained operators in the dual field theory. One can say that there is a passive and an active point of view of bulk reconstruction. In the active point of view advocated in the HKLL and the more refined JLMS procedures, we extrapolate the bulk operator to the boundary. On the other hand in the RG flow point of view, we should coarse-grain the boundary operator under a suitably defined RG flow such that it mimics the geometric radial flow, and then demonstrate the emergence of local bulk operators from these flowed operators. The coarse-graining is more specifically an evolution under a sequence of CP (completely positive) unital maps of the dual field theory operators.131313A CP unital map is the dual of a CPTP (completely positive trace preserving) map which evolves density matrices to density matrices. Here dual implies the map which takes us to the Heisenberg picture from the Schrodinger picture. The unital map preserves the identity. Interestingly, the real space RG has been already discussed explicitly from the quantum error correction perspective furuya2020 ; ghodrati2021_modFlow and the Petz map also naturally emerges in this context. The passive RG flow perspective also holds an enormous promise to reveal novel principles of bulk emergence as we discuss below. It is not a mere reinterpretation of the active (HKLL/JLMS) perspective since the latter is manifestly non-local while the RG flow tames non-locality in a controlled way.

Various proposals for the RG flow have been advocated Lee:2009ij ; Lee:2013dln ; Behr:2015aat ; Behr:2015yna ; Mandal:2016rmt ; Sathiapalan:2017frk based on fundamental insights developed in Heemskerk:2009pn ; Heemskerk:2010hk . Here we will focus on the highly efficient RG flow construction developed in Behr:2015aat ; Behr:2015yna based on earlier works in the context of the fluid/gravity correspondence Kuperstein:2011fn ; Kuperstein:2013hqa . For a review see Mukhopadhyay:2016fre . In this approach bulk locality is manifest especially through the Ward identities. We outline this approach here in the language of quantum error correction.

We first choose a code subspace. Let this be the space of all solutions of pure gravity with a negative cosmological constant. This class of states can be charactized by the expectation value of a specific single-trace operator, namely the energy-momentum tensor, i.e. tμνdelimited-⟨⟩subscript𝑡𝜇𝜈\langle t_{\mu\nu}\rangle since the expectation values of multi-trace operators factorize in the large N𝑁N limit and other single-trace operators have vanishing or fixed expectation values.141414To be precise one needs to also specify intial conditions to describe the full geometry. However, tμνdelimited-⟨⟩subscript𝑡𝜇𝜈\langle t_{\mu\nu}\rangle is all we need to find the geometry in a radial tube in Fefferman-Graham or Eddington-Finkelstein coordinates when the boundary metric (dual to the physical metric in which the dual field theory lives) is also specified. A better way to implement this would be to invoke a Borel resummation of the derivative expansion at late time in tμνdelimited-⟨⟩subscript𝑡𝜇𝜈\langle t_{\mu\nu}\rangle Heller:2013fn . Then the initial conditions are encoded in the Stokes parameters.

The aim is to define a sequence of CP unital maps parametrized by a coarse-graining scale ΛΛ\Lambda under which

tμνtμν(Λ).subscript𝑡𝜇𝜈subscript𝑡𝜇𝜈Λt_{\mu\nu}\rightarrow t_{\mu\nu}(\Lambda).

Since other single-trace operators do not play a role in this subspace, most generally we should obtain

tμν(Λ)subscript𝑡𝜇𝜈Λ\displaystyle t_{\mu\nu}(\Lambda) =\displaystyle= tμν+a11Λ2tμν+1Λ4(a2tμρtρν+a3ημνtαβtαβ+a42tμν+)subscript𝑡𝜇𝜈subscript𝑎11superscriptΛ2subscript𝑡𝜇𝜈1superscriptΛ4subscript𝑎2superscriptsubscript𝑡𝜇𝜌subscript𝑡𝜌𝜈subscript𝑎3subscript𝜂𝜇𝜈subscript𝑡𝛼𝛽superscript𝑡𝛼𝛽subscript𝑎4superscript2subscript𝑡𝜇𝜈\displaystyle t_{\mu\nu}+a_{1}\frac{1}{\Lambda^{2}}\Box t_{\mu\nu}+\frac{1}{\Lambda^{4}}(a_{2}t_{\mu}^{\,\,\rho}t_{\rho\nu}+a_{3}\eta_{\mu\nu}t_{\alpha\beta}t^{\alpha\beta}+a_{4}\Box^{2}t_{\mu\nu}+\cdots) (76)
+𝒪(1Λ6).𝒪1superscriptΛ6\displaystyle+\mathcal{O}\left(\frac{1}{\Lambda^{6}}\right).

Above aisubscript𝑎𝑖a_{i}s are appropriate numerical constants which are determined by the specific CP unital map (the insight that single trace operators should mix with multi-trace operators under the RG flow even in the large N𝑁N limit is from Heemskerk:2009pn ). Essentially we put in all possible multi-trace operators on the right hand side except those which vanish due to the CFT Ward identities:

μtμν=0,ηαβtαβ=0.formulae-sequencesuperscript𝜇subscript𝑡𝜇𝜈0subscript𝜂𝛼𝛽superscript𝑡𝛼𝛽0\partial^{\mu}t_{\mu\nu}=0,\quad\eta_{\alpha\beta}t^{\alpha\beta}=0. (77)

Owing to large-N𝑁N factorization of multi-trace operators, the evolution (76) is effectively a classical equation. The point of the highly efficient RG flow is that the coarse-graining which generates these unital maps should be done with a very specific choice of the complementary subspaces which are traced out such that we can define a metric

gμν(Λ)=gμν[tαβ(Λ)]subscript𝑔𝜇𝜈Λsubscript𝑔𝜇𝜈delimited-[]subscript𝑡𝛼𝛽Λg_{\mu\nu}(\Lambda)=g_{\mu\nu}[t_{\alpha\beta}(\Lambda)]

in which we obtain a local Ward identity

(Λ)μtμν(Λ)=0superscriptsubscriptΛ𝜇subscript𝑡𝜇𝜈Λ0\nabla_{(\Lambda)}^{\mu}t_{\mu\nu}(\Lambda)=0 (78)

at each scale ΛΛ\Lambda with (Λ)subscriptΛ\nabla_{(\Lambda)} being the covariant derivative constructed from gμν(Λ)subscript𝑔𝜇𝜈Λg_{\mu\nu}(\Lambda). The Ward identity is defined by considering tνμ(Λ)subscriptsuperscript𝑡𝜇𝜈Λt^{\mu}_{\,\,\nu}(\Lambda) as the fundamental variable and we lower and raise indices using gμν(Λ)subscript𝑔𝜇𝜈Λg_{\mu\nu}(\Lambda) and its inverse respectively. In practice, we need to then restrict aisubscript𝑎𝑖a_{i}s appearing in (76) such that we can obtain bisubscript𝑏𝑖b_{i}s defining gμν(Λ)subscript𝑔𝜇𝜈Λg_{\mu\nu}(\Lambda) via the expansion

gμν(Λ)subscript𝑔𝜇𝜈Λ\displaystyle g_{\mu\nu}(\Lambda) =\displaystyle= ημν+1Λ4tμν+b11Λ6tμν+subscript𝜂𝜇𝜈1superscriptΛ4subscript𝑡𝜇𝜈limit-fromsubscript𝑏11superscriptΛ6subscript𝑡𝜇𝜈\displaystyle\eta_{\mu\nu}+\frac{1}{\Lambda^{4}}t_{\mu\nu}+b_{1}\frac{1}{\Lambda^{6}}\Box t_{\mu\nu}+ (79)
1Λ8(b2tμρtρν+b3ημνtαβtαβ+b42tμν+)+𝒪(1Λ10)1superscriptΛ8subscript𝑏2superscriptsubscript𝑡𝜇𝜌subscript𝑡𝜌𝜈subscript𝑏3subscript𝜂𝜇𝜈subscript𝑡𝛼𝛽superscript𝑡𝛼𝛽subscript𝑏4superscript2subscript𝑡𝜇𝜈𝒪1superscriptΛ10\displaystyle\frac{1}{\Lambda^{8}}(b_{2}t_{\mu}^{\,\,\rho}t_{\rho\nu}+b_{3}\eta_{\mu\nu}t_{\alpha\beta}t^{\alpha\beta}+b_{4}\Box^{2}t_{\mu\nu}+\cdots)+\mathcal{O}\left(\frac{1}{\Lambda^{10}}\right)

with which (78) is satisfied. This is possible only for specific choices of aisubscript𝑎𝑖a_{i}s and hence the coarse-graining CP unital maps. The emergence of bulk spacetime follows by considering the D+1𝐷1D+1-dimensional metric in the Fefferman-Graham gauge after identifying r𝑟r with Λ1superscriptΛ1\Lambda^{-1}:

ds2=dr2+gμν(r,x)dxμdxνr2.dsuperscript𝑠2dsuperscript𝑟2subscript𝑔𝜇𝜈𝑟𝑥dsuperscript𝑥𝜇dsuperscript𝑥𝜈superscript𝑟2{\rm d}s^{2}=\frac{{\rm d}r^{2}+g_{\mu\nu}(r,x){\rm d}x^{\mu}{\rm d}x^{\nu}}{r^{2}}. (80)

The metric above satisfies the D+1𝐷1D+1-dimensional Einstein’s equations or appropriate classical gravity equations. We can always choose the aisubscript𝑎𝑖a_{i}s in (76) such that we obtain gμν(Λ=r1)subscript𝑔𝜇𝜈Λsuperscript𝑟1g_{\mu\nu}(\Lambda=r^{-1}) with which (78) is satisfied and the metric (80) satisfies D+1𝐷1D+1-dimensional classical gravity equations. It was shown in Behr:2015aat that (78) is sufficient to guarantee the emergence of a D+1𝐷1D+1-dimensional classical gravity because without the emergence of a gauge (diffeomorphism) symmetry (78) cannot hold along the RG flow. In fact the RG flow has an automorphism which is related to the residual gauge symmetry of the Fefferman-Graham gauge that maps to conformal transformations at the boundary. Via appropriate state-dependent conjugations one can obtain the RG flow that reproduces the bulk metric in other gauges. In the Fefferman-Graham gauge, the RG flow is manifestly state-independent (except for a subtlety which we describe below).

The key point of highly efficient RG flow is that one should choose the complementary subspace which is projected out such that the energy-momentum exchanges with the complement can be absorbed into a redefinition of the metric. One can generalize this to cases where other single-trace operators have non-trivial expectation values simply by introducing sources for these operators such that a general version of the local Ward identity (78) holds. These scale-dependent sources then are the dual bulk fields after we identify r𝑟r with Λ1superscriptΛ1\Lambda^{-1}.

This coarse-graining was implemented in the hydrodynamic sector in Behr:2015yna . To do this one assumes that tμνdelimited-⟨⟩subscript𝑡𝜇𝜈\langle t_{\mu\nu}\rangle is given in terms of constitutive relations via hydrodynamic variables which can be expanded in the derivative expansion. In this case, we need to sum to all orders in Λ1superscriptΛ1\Lambda^{-1} at a fixed order in derivatives. The RG flow is implemented by coarse-graining of the hydrodynamic variables with the scale itself being a local functional of these variables (see Fig. 15). In this case, the RG flow (76) reduces simply to first order ODEs which evolve infinite number of transport coefficients. This reproduces Einstein’s equations (or other classical gravity equations). To see this we simply need to follow the procedure in Kuperstein:2011fn ; Kuperstein:2013hqa and rewrite the gravity equations (in the derivative expansion) as a first order flow of transport coefficients.

Refer to caption
Figure 15: The highly efficient RG flow in the hydrodynamic sector can be simply constructed as evolution of the velocity field uμsuperscript𝑢𝜇u^{\mu} and the temperature field T𝑇T with the coarse-graining length scale (identified with the bulk radial coordinate r𝑟r) via the constitutive relation defining the stress-tensor. The effective background metric is evolved such that the stress tensor is locally conserved. This implies that the scale of coarse-graining ΛΛ\Lambda becomes dependent on uμ(r)superscript𝑢𝜇𝑟u^{\mu}(r) and T(r)𝑇𝑟T(r) in a very specific way so that the violation of the local energy-momentum conservation can be absorbed into a re-definition of the metric. The highly efficient RG flow simply reduces to the first order flow equations for the transport coefficients. These have unique solutions requiring that at the end-point corresponding to the location of the horizon we obtain an incompressible non-relativistic Navier-Stokes fixed point. This leads to values of transport coefficients at the boundary which give a dual spacetime without naked singularities at the future horizon. Figure from Kuperstein:2013hqa .

This raises a profound question. In the fluid/gravity correspondence Bhattacharyya:2007vjd ; Baier:2007ix , one explicitly solves the equations of gravity in derivative expansion and obtains the transport coefficients in the UV by requiring that the dual solution has a regular future horizon. However in the RG flow we do not see the metric directly but only tμν(Λ)subscript𝑡𝜇𝜈Λt_{\mu\nu}(\Lambda). How do we then determine the transport coefficients? This was solved in Kuperstein:2013hqa . It was shown that there is a particular scale ΛhsubscriptΛ\Lambda_{h} where there is naively a singularity in the RG flow with the pressure and most transport coefficients blowing up – this scale corresponds to the horizon in the dual geometry. However, if one redefines time and the scale in a universal way (analogous to taking the near-horizon limit), then the singularity actually transforms to a fixed point described by non-relativistic incompressible Navier-Stokes equations with a single parameter, namely the shear-viscosity. This is however possible only if the transport coefficients do not blow up faster than certain bounds as ΛΛ\Lambda approaches ΛhsubscriptΛ\Lambda_{h}. These conditions then fix the integration constants in the ODEs describing the RG flow of the transport coefficients, and we precisely obtain those values of the transport coefficients (shear-viscosity at first order and five other transport coefficients at second order etc) at the boundary computed in Policastro:2001yc ; Bhattacharyya:2007vjd ; Baier:2007ix which lead to a regular future horizon. Essentially the regularity of spacetime is simply a consequence of the regularity of the highly efficient RG flow which in the hydrodynamic limit is simply the condition that it ends at the scale of the mean-free path at a fixed point described by the non-relativistic incompressible Navier-Stokes equations (governed just by an effective shear viscosity).151515In Kuperstein:2013hqa , it was shown that the counterterms which remove the UV divergences are also fixed by the Navier-Stokes fixed point.

The highly efficient RG flow can lead to a new way to understand bulk emergence. The RG flow can be naturally viewed as a encoding of infrared (coarse-grained) physics into microscopic degrees of freedom. The highly efficient RG flow principle of preservation of effective Ward identities (78) then leads to emergence of bulk fields. It might seem that at least in the Fefferman-Graham gauge this RG flow (76) is manifestly state-independent. However, this is not completely true because one needs to impose infrared boundary conditions which depends on the code subspace. In the case of the hydrodynamic sector this is specified by the Navier-Stokes fixed point. However, within the code subspace the RG flow can be cast in a state-independent form. This is precisely how the reconstruction of the black hole interior in the dual field theory should be state-dependnet as we will discuss in section 5.2. The boundary operator which will reconstruct the bulk operator in the black hole interior will depend on the choice of the code subspace but not on the specific state in this code subspace. This motivates a study of the highly efficient RG flow beyond the hydrodynamic sector.

5 Decoding the black hole interior

5.1 How the islands emerge from replica wormholes

The black hole interior poses the most formidable challenge in the understanding of bulk reconstruction. If the evolution of the black hole can be described by an unitary theory, we should expect that after the Page time (defined below), the information of the interior should begin to leak out into the Hawking radiation. This is a consequence of Page’s theorem Page_1993 ; Page_2013 which states that for a typical pure state in a bipartite system with one system being much smaller than the other, the trace distance between the density matrix of the smaller subsystem and the microcanonical ensemble is of the order of eS/2superscript𝑒𝑆2e^{-S/2} with S𝑆S being the number of qubits in the larger one. At initial stages, the black hole is much larger than its Hawking radiation, so the entanglement entropy of the latter should grow. The Page time occurs when the black hole and the emitted radiation have the same coarse grained entropy161616The coarse-grained entropy is defined as follows. Consider expectations values of simple operators (averaged over certain time-scales) and construct all density matrices which reproduce them. The coarse-grained entropy is the entropy of such a density matrix which has maximal von-Neumann entropy. Since the argument of Page uses typicality, the Page time is appropriately defined via the coarse-grained entropy which is determined mostly by the size of the Hilbert space.. After the Page time, the entanglement entropy of the radiation should decrease as it becomes the larger subsystem. The already emitted Hawking quanta should also be purified by the subsequently emitted Hawking quanta. Thus its total entanglement entropy should follow the Page curve – initially it should grow linearly in time following Hawking’s original computation but then it should fall back to zero after Page time as it gets purified.

The crucial question is that if the semi-classical description is valid for the effective field theory (EFT) observables, then what exactly goes wrong with the semi-classical computation of the entanglement entropy of the Hawking radiation after the Page time. The question is further sharpened by the Almheiri-Marolf-Polchinski-Sully (AMPS) Almheiri:2012rt paradox which points that such assumptions would lead to violation of the monogamy of entanglement because the Hawking quanta are maximally entangled with their infalling counterparts according to the EFT computations, but they also have to be maximally entangled with the quanta emitted before Page time in order to purify it and validate the unitarity of the evolution. The strong sub-additivity property of the entanglement entropy would prevent simultaneous maximal entanglement of a system with two other systems PhysRevA.69.022309 . We will discuss more aspects of this paradox later.171717The AMPS paradox builds on the discussions by Mathur Mathur:2009hf which introduced the strong subadditivity originally to examine if small semi-classical corrections can restore unitarity in Hawking radiation. Many aspects of the AMPS paradox were discussed also in PhysRevLett.110.101301 .

It is to this question (paradox) that the AdS/CFT correspondence has recently produced some remarkable new insights along with a deeper understanding of which part of the black hole interior is encoded in a given subregion of the Hawking radiation system. The main lesson is that the EFT in the semi-classical black hole geometry is indeed unimpeachable in its usual domain of validity set by an energy scale, however there are other subleading Euclidean saddles in the path integral for Rènyi entropies which give sufficient contribution after Page time to restore unitarity. Furthermore, these saddles imply that the fine-grained entropy181818The fine-grained entropy can be defined like the coarse-grained entropy above but taking into account more operators that probe the short distance structure of the density matrix. However, we should keep in mind that we are already doing an averaging which allows a tensor factorization between the black hole and radiation system to emerge. A very concrete discussion on this issue can be found in Ghosh:2021axl within the context of the semi-classical approximation itself. We will have a more elaborate discussion on this in the context of microstate models in Section 5.3.2. of Hawking radiation (obtained from the Rènyi entropies) would involve contributions from islands (which are regions of bulk spacetime diconnected from the asymptotic boundary and including portions of the black hole interior as already introduced in the context of explicit models in section 3.5) after Page time as these are connected to the already emitted quanta via wormholes. This implies the ER= EPR mechanism Maldacena_2013 of resolution of AMPS paradox with wormholes implying that after Page time operators in the interior also affect this far away radiation, and therefore the early radiation and the interior are not separable systems. We will discuss later that if one has to resolve this issue by explicitly following the state in real time we will need self-averaging and complexity. The crucial inputs in both cases would be the island reconstruction for which the framework of operator error correction would play an important role.

Refer to caption
Figure 16: This figure is from Almheiri2020Islands . It depicts the construction where a holographic system B𝐵B is coupled to a large system R𝑅R without gravity. The dynamics of B𝐵B is described by the evaporating black hole with a constant curvature which is glued to flat Minkowski space in which R𝑅R lives. R𝑅R collects the Hawking quanta of the evaporating black hole. The red regions are the entanglement wedges of the interval(s) which contribute to the von-Neumann entropy of the Hawking quanta in R𝑅R while the green regions are those that contribute to B𝐵B. (a) At early time, the entropy of the matter fields in whole of the Cauchy slice contributes to B𝐵B. (b) After Page time, the interval exterior to the quantum extremal surface, whose entanglement wedge is the island, contributes to the von-Neumann entropy of the Hawking quanta in R𝑅R.

The anti-de Sitter black hole cannot evaporate unless the boundary conditions allow the quanta of bulk matter to escape out of the asymptotic region. The understanding of how the black hole interior is encoded into these Hawking quanta especially after the Page time can be achieved by coupling a holographic system B𝐵B with another larger system R𝑅R living in one higher dimension, and which may or may not be holographic. The role of R𝑅R is to collect the Hawking quanta of the evaporating black hole in the dual geometry that depicts the evolving system B𝐵B holographically. For tractability, we also need the bulk matter to have large central charge c𝑐c so that we can ignore quantum gravity fluctuations. What has been shown to be crucial is that, after the Page time, the entanglement wedge of the quantum extremal surface (QES) of the asymptotic boundary is encoded in B𝐵B, while the wedge including the exterior of the QES called the island and containing parts of the black hole interior is encoded into R𝑅R as illustrated in Fig. 16 for 0+1010+1-dimensional B𝐵B coupled to a 1+1111+1-dimensional R𝑅R. To see this, we need to consider the (pure) state of the quantum matter fields on a Cauchy slice of the black hole geometry and R𝑅R, and consider contributions of saddles that contribute to the Rènyi entropy of the Hawking radiation in R𝑅R other than the semi-classical black hole itself. These saddles connect the replicas of the Rènyi entropy computation via wormholes, and are therefore called replica wormholes. These wormholes connect the island I𝐼I including parts of the black hole interior to the Hawking quanta in R𝑅R, thus invalidating the tri-partition in the AMPS paradox and resolving it.

Refer to caption
Figure 17: A thermofield double version of the setup described in Fig. 16 taken from Almheiri2020Islands . The particles with the same color (blue/green) are entangled according to effective field theory computations. Here we consider the entanglement entropy of the Hawking quanta in the red intervals at time t=0𝑡0t=0 and at t=t0𝑡subscript𝑡0t=t_{0} with t0subscript𝑡0t_{0} exceeding the Page time. In (a) we see that if we do not include the island, then the von-Neumann entropy of the red intervals (R𝑅R )should grow linearly with time and with the slope determined by the rate of Hawking pair production. We see in (b) that if one includes the island (I𝐼I), then the matter contribution to the von-Neumann entropy in IR𝐼𝑅I\cup R cannot grow as the entangled pairs get collected in this combined region.

We need to explain why such saddles which give exponentially small contributions in the semi-classical limit to the Rènyi entropy should be important at all. The leading saddle is the usual ~n/Znsubscript~𝑛subscript𝑍𝑛\widetilde{\mathcal{M}}_{n}/Z_{n} geometry discussed in Section 2 denoting n𝑛n copies of the semi-classical black hole geometry quotiented by the cyclic permutation symmetry. In the presence of gravity, we need to take into account the other replica wormhole saddles too. The point is that the contribution from the ~n/Znsubscript~𝑛subscript𝑍𝑛\widetilde{\mathcal{M}}_{n}/Z_{n} saddle decreases exponentially with time (consistent with the linear growth of the von-Neumann entropy in the R𝑅R region) while the contribution from the other replica wormhole saddles grow with time essentially due to the growth of the island. Around the Page time, these effects compete with each other, and after sufficiently long time the replica wormholes give the leading contribution to the Rènyi entropies, and arrest the growth of the von-Neumann entropy of R𝑅R as illustrated in Fig. 17.191919Note in the thermofield double setup the entanglement entropy of the radiation should not vanish but rather saturate to be consistent with unitarity since not all radiation escapes to the null infinities in R𝑅R. The Page curve then involves a transition from linear growth to saturation.

Refer to caption
Figure 18: An example of a replica wormhole from Almheiri2020Islands for the computation of the 333rd order Rènyi entropy of the Hawking quanta in R𝑅R in the setup described in Fig. 17. Firstly, we consider three copies of the Euclidean version of the diagrams in Fig. 17. In each copy, the shaded blue region is the Poincaré disk denoting Euclidean AdS2𝐴𝑑subscript𝑆2AdS_{2} black hole with an appropriate time-reparametrization at the boundary of the disk. The white regions have cuts corresponding to the red intervals in Fig. 17. In (a) we denote the covering manifold ~3subscript~3\widetilde{\mathcal{M}}_{3} and in (b) the actual replica wormhole solution which is 3=~3/Z3subscript3subscript~3subscript𝑍3\mathcal{M}_{3}=\widetilde{\mathcal{M}}_{3}/Z_{3} which arises as a solution to the Euclidean gravity coupled to matter. The fixed points of the Z3subscript𝑍3Z_{3} action, namely w1subscript𝑤1w_{1} and w2subscript𝑤2w_{2}, appear as conical singularities in the gravitational solution where the bulk twist operators are inserted. Note their positions are determined by the equations of motion also.

These replica wormholes are essentially codimension two cosmic branes on the gravity side (giving rise to conical singularities in two-dimensional gravity) where the twist operators of the bulk matter theory must be inserted. See Fig. 18 for an illustration. The positions of the these branes should be obtained by solving the gravitational equations of motion. For the semi-infinite interval in R𝑅R containing the Hawking quanta in Fig. 16, it turns out that there is one such brane, and in the limit n1𝑛1n\rightarrow 1 with n𝑛n denoting the number of replicas, the position of the cosmic brane is exactly that of the quantum extremal surface which computes the entropy of B𝐵B. In case of the double interval in Fig. 17, the cosmic branes coincide with the two quantum extremal surfaces in the limit n1𝑛1n\rightarrow 1 as well. For a recent pedagogical review on wormholes in semiclassical gravity including the explicit construction of these replica wormholes see Arnab_review .

Specifically, the island rule (54) which is now validated by the gravitational path integral accounting for the replica wormhole saddles, implies that

S(ρR)=S(ρRI(g))+A(QES)4G.𝑆subscript𝜌𝑅𝑆superscriptsubscript𝜌𝑅𝐼𝑔𝐴QES4𝐺S(\rho_{R})=S(\rho_{RI}^{(g)})+\frac{A({\rm QES})}{4G}. (81)

Above ρRsubscript𝜌𝑅\rho_{R} denotes the density matrix of the interval(s) in R𝑅R, ρRI(g)superscriptsubscript𝜌𝑅𝐼𝑔\rho_{RI}^{(g)} denotes the density matrix of the quantum fields (responsible for the Hawking quanta) in RI𝑅𝐼R\cup I in the semiclassical geometry of the black hole glued to Minkowski space(s). Note that A(QES)𝐴QESA({\rm QES}) above refers simply to the value of the dilaton for the JT theory since the QES is a point (or two points in the case of the thermofield double). We should understand the above as a semi-classical statement only – the island is defined with respect to the semi-classical geometry itself. On the other hand, if D𝐷D is the interval in this geometry which is the complement of RI𝑅𝐼R\cup I and connects the QES to the boundary of the AdS2𝐴𝑑subscript𝑆2AdS_{2} region, then the generalization of the HRT formula given by (53) will imply that

S(ρB)=S(ρD(g))+A(QES)4G𝑆subscript𝜌𝐵𝑆superscriptsubscript𝜌𝐷𝑔𝐴QES4𝐺S(\rho_{B})=S(\rho_{D}^{(g)})+\frac{A({\rm QES})}{4G} (82)

Above by B𝐵B we mean its two copies in the thermofield double case. Since, the quantum fields in the full Cauchy slice IDR𝐼𝐷𝑅I\cup D\cup R in the semiclassical description should be in a pure state, it implies that S(ρRI(g))=S(ρD(g))𝑆superscriptsubscript𝜌𝑅𝐼𝑔𝑆superscriptsubscript𝜌𝐷𝑔S(\rho_{RI}^{(g)})=S(\rho_{D}^{(g)}), and therefore S(ρR)=S(ρB)𝑆subscript𝜌𝑅𝑆subscript𝜌𝐵S(\rho_{R})=S(\rho_{B}). This is then consistent with unitarity of the evolution of the full (two copies of the) dual BR𝐵𝑅B\cup R system.

We can also then claim that the bulk effective field theory operators acting in the island can be reconstructed in R𝑅R as an extension of the entanglement wedge reconstruction described before in section 3. The code subspace is the span of states in the full (two copies of) RB𝑅𝐵R\cup B which can be constructed in the gravitational description via operator insertions in the same semi-classical geometry containing the same island Iisubscript𝐼𝑖I_{i} at leading order in perturbative EFT corresponding to an interval Risubscript𝑅𝑖R_{i} in R𝑅R. The code subspace is assumed to admit a factorization of the form code=RiR¯isubscript𝑐𝑜𝑑𝑒tensor-productsubscriptsubscript𝑅𝑖subscriptsubscript¯𝑅𝑖\mathcal{H}_{code}=\mathcal{H}_{R_{i}}\otimes\mathcal{H}_{\overline{R}_{i}} where Risubscript𝑅𝑖R_{i} here denotes the interval(s) in R𝑅R, and R¯isubscript¯𝑅𝑖\overline{R}_{i} the complement of Risubscript𝑅𝑖R_{i} that includes (two copies of) B𝐵B. The EFT of bulk fields in the semi-classical gravity description including the Minkowski space region(s) imply that code=RiDiIisubscript𝑐𝑜𝑑𝑒tensor-productsubscriptsubscript𝑅𝑖subscriptsubscript𝐷𝑖subscriptsubscript𝐼𝑖\mathcal{H}_{code}=\mathcal{H}_{R_{i}}\otimes\mathcal{H}_{D_{i}}\otimes\mathcal{H}_{I_{i}} where Disubscript𝐷𝑖D_{i} is the complement of the island and Risubscript𝑅𝑖R_{i}. Then (81) holds for any ρRiksuperscriptsubscript𝜌subscript𝑅𝑖𝑘\rho_{R_{i}}^{k} obtained from any state |kcodeket𝑘subscript𝑐𝑜𝑑𝑒|k\rangle\in\mathcal{H}_{code} by tracing out R¯isubscript¯𝑅𝑖\overline{R}_{i}, and with corresponding ρRiIi(g)ksuperscriptsubscript𝜌subscript𝑅𝑖subscript𝐼𝑖𝑔𝑘\rho_{R_{i}I_{i}}^{(g)k} on the bulk side obtained by tracing out Disubscript𝐷𝑖D_{i}. An extension of the JLMS argument would then imply that the relative entropies of the semi-classical bulk states and the corresponding states in the full dual description are identical, i.e.

S(ρRim|ρRin)=S(ρRiIi(g)m|ρRiIi(g)n)𝑆conditionalsuperscriptsubscript𝜌subscript𝑅𝑖𝑚superscriptsubscript𝜌subscript𝑅𝑖𝑛𝑆conditionalsuperscriptsubscript𝜌subscript𝑅𝑖subscript𝐼𝑖𝑔𝑚superscriptsubscript𝜌subscript𝑅𝑖subscript𝐼𝑖𝑔𝑛S(\rho_{R_{i}}^{m}|\rho_{R_{i}}^{n})=S(\rho_{R_{i}I_{i}}^{(g)m}|\rho_{R_{i}I_{i}}^{(g)n}) (83)

for two states |mket𝑚|m\rangle and |nket𝑛|n\rangle belonging to codesubscript𝑐𝑜𝑑𝑒\mathcal{H}_{code}. A repeat of the theorem of Dong, Harlow and Wall described in section 4.2.1 then implies that for any bulk EFT operator 𝒪Iisubscript𝒪subscript𝐼𝑖\mathcal{O}_{I_{i}} localized in the island Iisubscript𝐼𝑖I_{i} and any |ncodeket𝑛subscript𝑐𝑜𝑑𝑒|n\rangle\in\mathcal{H}_{code}, we should have

  1. 1.

    For any any X𝑋X localized in R¯isubscript¯𝑅𝑖\overline{R}_{i}

    n|[X,OIi]|n=0,quantum-operator-product𝑛𝑋subscript𝑂subscript𝐼𝑖𝑛0\langle n|[X,O_{I_{i}}]|n\rangle=0, (84)

    and

  2. 2.

    there exists an operator 𝒪Risubscript𝒪subscript𝑅𝑖\mathcal{O}_{R_{i}} localized in Risubscript𝑅𝑖R_{i} such that its action on the code subspace is identical, i.e.

    𝒪Ri|n=𝒪Ii|n,𝒪Ri|n=𝒪Ii|n.formulae-sequencesubscript𝒪subscript𝑅𝑖ket𝑛subscript𝒪subscript𝐼𝑖ket𝑛superscriptsubscript𝒪subscript𝑅𝑖ket𝑛superscriptsubscript𝒪subscript𝐼𝑖ket𝑛\mathcal{O}_{R_{i}}|n\rangle=\mathcal{O}_{I_{i}}|n\rangle,\quad\mathcal{O}_{R_{i}}^{\dagger}|n\rangle=\mathcal{O}_{I_{i}}^{\dagger}|n\rangle. (85)

Although the replica wormhole saddles produce results which are compatible with unitarity, there are crucial subtleties which are explicitly brought out by the Euclidean (Penington, Shenker, Stanford and Yang) PSSY toy model Penington:2019kki in which we can sum over the planar replica wormhole saddles explicitly (via a Schwinger-Dyson type equation for the resolvent matrix 1/(λIρR)1𝜆𝐼subscript𝜌𝑅1/(\lambda I-\rho_{R})). It also gives a simple way to see how the replica wormhole saddles give rise to the island which we will review below. In this Euclidean model, we represent the state of the BR𝐵𝑅B\cup R system in the maximally entangled form

|ψ=1ki=1k|ψiB|iRket𝜓1𝑘superscriptsubscript𝑖1𝑘subscriptketsubscript𝜓𝑖𝐵subscriptket𝑖𝑅|\psi\rangle=\frac{1}{\sqrt{k}}\sum_{i=1}^{k}|\psi_{i}\rangle_{B}|i\rangle_{R} (86)

which should be a good approximation to the full state of the Hawking quanta especially after Page time. The state of the holographic bulk B𝐵B is captured by an end of the world (EOW) brane state i𝑖i as shown in Fig. 19. The evolution of time is captured by simply dialling k𝑘k. To capture the dynamics post Page time then we need keSBHmuch-greater-than𝑘superscript𝑒subscript𝑆𝐵𝐻k\gg e^{S_{BH}} where SBHsubscript𝑆𝐵𝐻S_{BH} is the entropy of the bulk black hole. The bulk theory is pure JT gravity with a dilaton and no bulk matter, while the EOW brane has vanishing extrinsic curvature and the Neumann boundary condition for the bulk dilaton is also imposed at its location. Explicitly, the Euclidean action for the metric g𝑔g and the dilation ϕitalic-ϕ\phi is

S𝑆\displaystyle S =\displaystyle= S04π(gR+2hK)subscript𝑆04𝜋subscript𝑔𝑅2subscript𝐾\displaystyle-\frac{S_{0}}{4\pi}\left(\int_{\mathcal{M}}\sqrt{g}R+2\int_{\partial\mathcal{M}}\sqrt{h}K\right) (87)
12(gϕ(R+2)+2hϕK)12subscript𝑔italic-ϕ𝑅22subscriptitalic-ϕ𝐾\displaystyle-\frac{1}{2}\left(\int_{\mathcal{M}}\sqrt{g}\phi(R+2)+2\int_{\partial\mathcal{M}}\sqrt{h}\phi K\right)
+μEOWbraneds,𝜇subscript𝐸𝑂𝑊𝑏𝑟𝑎𝑛𝑒differential-d𝑠\displaystyle+\mu\int_{EOW-brane}{\rm d}s,

with hh denoting the induced metric on the boundary and μ0𝜇0\mu\geq 0 the mass of the EOW brane. The mentioned boundary conditions imply that on the EOW brane we should impose

K=0,nϕ=μformulae-sequence𝐾0subscript𝑛italic-ϕ𝜇K=0,\quad\partial_{n}\phi=\mu (88)

with nsubscript𝑛\partial_{n} denoting the normal derivative. The standard asymptotic boundary conditions imply that at the boundary

h=1ϵ2,ϕ=1ϵ,ϵ0.formulae-sequence1superscriptitalic-ϵ2formulae-sequenceitalic-ϕ1italic-ϵitalic-ϵ0h=\frac{1}{\epsilon^{2}},\quad\phi=\frac{1}{\epsilon},\quad\epsilon\rightarrow 0. (89)

The limit ϵ0italic-ϵ0\epsilon\rightarrow 0 implies implementing the standard procedure of holographic renormalization which extracts physical observables of the B𝐵B system from the regularized action of the classical gravity solution. The solutions of the gravitational theory are simply the black hole solutions. It is easy to note that the computations of the replica wormhole saddles simplify in absence of bulk matter too. The results of the computations are the same. When keSBHmuch-greater-than𝑘superscript𝑒subscript𝑆𝐵𝐻k\gg e^{S_{BH}}, the replica wormhole saddles produce a dominant contribution as a result of which the QES in the limit n1𝑛1n\rightarrow 1 is close to the bifurcate horizon opening up an island which is the entanglement wedge of the R𝑅R system.

Refer to caption
Figure 19: The simplified PSSY model with an end of the world brane carrying a degree of freedom. It has zero extrinsic curvature and the Neumann boundary condition for the JT gravity dilaton is imposed on it. The left and right figures correspond to the Euclidean and Lorentzian scenarios. Figure from Penington:2019kki

The subtlety is that we would naively think that if the EOW brane states |iket𝑖|i\rangle are orthogonal, then the reduced of the R𝑅R system which follows from (86) is

ρR=1ki=1k|ii|R.subscript𝜌𝑅1𝑘superscriptsubscript𝑖1𝑘ket𝑖subscriptbra𝑖𝑅\rho_{R}=\frac{1}{k}\sum_{i=1}^{k}|i\rangle\langle i|_{R}. (90)

Furthermore, Tr(ρR2)=1/kTrsuperscriptsubscript𝜌𝑅21𝑘{\rm Tr}(\rho_{R}^{2})=1/k. The latter is however not correct if we take into account the replica wormhole saddle as illustrated in Fig. 20. Explicitly we obtain

Tr(ρR2)=kZ(1)2+k2Z(2)(kZ(1))2=1k+Z(2)Z(1)2Trsuperscriptsubscript𝜌𝑅2𝑘superscriptsubscript𝑍12superscript𝑘2subscript𝑍2superscript𝑘subscript𝑍121𝑘subscript𝑍2superscriptsubscript𝑍12{\rm Tr}(\rho_{R}^{2})=\frac{kZ_{(1)}^{2}+k^{2}Z_{(2)}}{(kZ_{(1)})^{2}}=\frac{1}{k}+\frac{Z_{(2)}}{Z_{(1)}^{2}} (91)

where Z(2)subscript𝑍2Z_{(2)} is the partition function of the two-sided wormhole and Z(1)subscript𝑍1Z_{(1)} is the disk partition function (the denominator comes from the appropriate normalization of the path integral). This is clearly incompatible with (90).

Refer to caption
Figure 20: The computation of Tr(ρR2)Trsuperscriptsubscript𝜌𝑅2{\rm Tr}(\rho_{R}^{2}) involves two saddles. The first one is the product of two disconnected disks and the second one is a two-sided replica wormhole. To compute the trace we must join the lines dashed lines. The disconnected disks then produce a single loop implying a factor of k𝑘k while the double sided wormhole produces two loops implying a factor of k2superscript𝑘2k^{2}. Thus we obtain the numerator in Eq. (91). Figure from Penington:2019kki

The reconciliation is achieved if we interpret that the replica wormhole saddles of JT gravity are essentially computing the path integrals for an ensemble average of Hamiltonians describing the B𝐵B system, so that

ρR=1ki,j=1k|ij|RΨj|Ψi¯Bsubscript𝜌𝑅1𝑘superscriptsubscript𝑖𝑗1𝑘ket𝑖subscriptbra𝑗𝑅subscript¯inner-productsubscriptΨ𝑗subscriptΨ𝑖𝐵\rho_{R}=\frac{1}{k}\sum_{i,j=1}^{k}|i\rangle\langle j|_{R}\overline{\langle\Psi_{j}|\Psi_{i}\rangle}_{B} (92)

and

Tr(ρR2)=1k2i,j=1k|Ψj|Ψi|2¯B.Trsuperscriptsubscript𝜌𝑅21superscript𝑘2superscriptsubscript𝑖𝑗1𝑘subscript¯superscriptinner-productsubscriptΨ𝑗subscriptΨ𝑖2𝐵{\rm Tr}(\rho_{R}^{2})=\frac{1}{k^{2}}\sum_{i,j=1}^{k}\overline{|\langle\Psi_{j}|\Psi_{i}\rangle|^{2}}_{B}. (93)

Supposing that

Ψj|ΨiB=δij+eS0/2Rijsubscriptinner-productsubscriptΨ𝑗subscriptΨ𝑖𝐵subscript𝛿𝑖𝑗superscript𝑒subscript𝑆02subscript𝑅𝑖𝑗\langle\Psi_{j}|\Psi_{i}\rangle_{B}=\delta_{ij}+e^{-S_{0}/2}R_{ij} (94)

where S0subscript𝑆0S_{0} is the entropy of the JT gravity black hole at zero temperature and Rijsubscript𝑅𝑖𝑗R_{ij} are random phases, we should obtain

Ψj|Ψi¯B=δij,|Ψj|Ψi|2¯B=δij+𝒪(eS0)formulae-sequencesubscript¯inner-productsubscriptΨ𝑗subscriptΨ𝑖𝐵subscript𝛿𝑖𝑗subscript¯superscriptinner-productsubscriptΨ𝑗subscriptΨ𝑖2𝐵subscript𝛿𝑖𝑗𝒪superscript𝑒subscript𝑆0\overline{\langle\Psi_{j}|\Psi_{i}\rangle}_{B}=\delta_{ij},\quad\overline{|\langle\Psi_{j}|\Psi_{i}\rangle|^{2}}_{B}=\delta_{ij}+\mathcal{O}(e^{-S_{0}}) (95)

where the bar on top denotes averaging over theories. Note that eS0superscript𝑒subscript𝑆0e^{-S_{0}} is exactly the order of magnitude of Z(2)/Z(1)2subscript𝑍2superscriptsubscript𝑍12Z_{(2)}/Z_{(1)}^{2} because of the respective topologies (two discs vs one with each disk being 𝒪(eS0)𝒪superscript𝑒subscript𝑆0\mathcal{O}(e^{S_{0}})202020The first term in the action (87) is topological and accounts for eS0superscript𝑒subscript𝑆0e^{S_{0}} weighting of the disk.). Note it is easy to see from (91) that when keS0much-greater-than𝑘superscript𝑒subscript𝑆0k\gg e^{S_{0}}, then the replica wormhole gives the dominant contribution Z(2)/Z(1)2subscript𝑍2superscriptsubscript𝑍12Z_{(2)}/Z_{(1)}^{2}. This in fact generalizes, and we can readily see that

Tr(ρRn)knZ(n)(kZ(1))n=Z(n)Z(1)n=𝒪(e(n1)S0)Trsuperscriptsubscript𝜌𝑅𝑛superscript𝑘𝑛subscript𝑍𝑛superscript𝑘subscript𝑍1𝑛subscript𝑍𝑛superscriptsubscript𝑍1𝑛𝒪superscript𝑒𝑛1subscript𝑆0{\rm Tr}(\rho_{R}^{n})\approx\frac{k^{n}Z_{(n)}}{(kZ_{(1)})^{n}}=\frac{Z_{(n)}}{Z_{(1)}^{n}}=\mathcal{O}(e^{-(n-1)S_{0}}) (96)

so that the contribution is entirely from the simply connected replica wormhole with n𝑛n asymptotic boundaries (see Fig. 21 for an illustration in the case of n=6𝑛6n=6). The Znsubscript𝑍𝑛Z_{n} quotient of these wormholes Z(n)subscript𝑍𝑛Z_{(n)} can be readily analytically continued (the Znsubscript𝑍𝑛Z_{n} symmetry is simply a rotational symmetry and quotienting produces a fixed point which is exactly the horizon in the limit n1𝑛1n\rightarrow 1 in which the geometry reduces to the original unreplicated one). It is also then easy to see that in the limit n1𝑛1n\rightarrow 1, the von Neumann entropy is just the generalized entropy of the QES that is located at the horizon, i.e.

Tr(ρR)=S0+2πϕh=SBH,Trsubscript𝜌𝑅subscript𝑆02𝜋subscriptitalic-ϕsubscript𝑆𝐵𝐻{\rm Tr}(\rho_{R})=S_{0}+2\pi\phi_{h}=S_{BH},

where ϕhsubscriptitalic-ϕ\phi_{h} is the value of the dilaton at the horizon and SBHsubscript𝑆𝐵𝐻S_{BH} is the thermodynamic entropy of the black hole. The latter equality follows from the action (87)) and note that there is no bulk matter in this model. Thus the k𝑘k\rightarrow\infty limit indeed implies that the QES is at the bifurcate horizon in the Lorentzian picture and the island is the full interior of the black hole. We explicitly see that the replica wormholes reproduce the correct QES. In the opposite limit kϵ𝑘italic-ϵk\rightarrow\epsilon with fixed S0subscript𝑆0S_{0}, the fully disconnected geometry dominates (see Fig. 21 for an illustration in the case of n=6𝑛6n=6) so that

Tr(ρRn)kZ(1)n(kZ(1))n=1kn1.Trsuperscriptsubscript𝜌𝑅𝑛𝑘superscriptsubscript𝑍1𝑛superscript𝑘subscript𝑍1𝑛1superscript𝑘𝑛1{\rm Tr}(\rho_{R}^{n})\approx\frac{kZ_{(1)}^{n}}{(kZ_{(1)})^{n}}=\frac{1}{k^{n-1}}. (97)

The von-Neumann entropy is then

Tr(ρR)=logkTrsubscript𝜌𝑅𝑘{\rm Tr}(\rho_{R})=\log k

in this limit signifying the absence of any island and that we should trace over the entire black hole spacetime which thus forms the entanglement wedge of B𝐵B. This toy model then reproduces features of the growth of the islands in the full Lorentzian computation with the bulk matter fields described earlier. The island vanishes at early time, and as the black hole evaporates it forms and eventually encompass the original interior. The flow of time is indeed represented in the toy model as the growth of k𝑘k as illustrated in Fig. 22.

Refer to caption
Figure 21: The computation of Tr(ρR6)Trsuperscriptsubscript𝜌𝑅6{\rm Tr}(\rho_{R}^{6}) involves several saddles of which the fully disconnected and the fully connected ones are explicitly shown above. The first one dominates when keSBHmuch-less-than𝑘superscript𝑒subscript𝑆𝐵𝐻k\ll e^{S_{BH}} while the second one dominates when keSBHmuch-greater-than𝑘superscript𝑒subscript𝑆𝐵𝐻k\gg e^{S_{BH}}. The latter has a Z6subscript𝑍6Z_{6} rotational symmetry whose action has a fixed point. Figure from Penington:2019kki .

The toy model also allows us to derive a version of the Petz map via path integrals with replica wormhole saddles for reconstruction of the operators acting on the island in the R𝑅R system. This makes sense if the wormholes perform an ensemble averaging over theories describing the B𝐵B system. The replica wormholes dominate when kdcodeeS0much-greater-than𝑘subscript𝑑codesuperscript𝑒subscript𝑆0k\gg d_{\rm code}e^{S_{0}} where dcodesubscript𝑑coded_{\rm code} is the dimension of the code subspace, and in this limit the mentioned Petz map produces a perfect reconstruction as expected. Of course when kdcodeeS0much-less-than𝑘subscript𝑑codesuperscript𝑒subscript𝑆0k\ll d_{\rm code}e^{S_{0}}, there is no island and the R𝑅R system should have no knowledge of the black hole interior.

Refer to caption
Figure 22: As described in the text, the fully disconnected replica wormholes dominate in the limit keSBHmuch-less-than𝑘superscript𝑒subscript𝑆𝐵𝐻k\ll e^{S_{BH}} implying that Tr(ρRn)=1/kn1Trsuperscriptsubscript𝜌𝑅𝑛1superscript𝑘𝑛1{\rm Tr}(\rho_{R}^{n})=1/k^{n-1}. This is obtained by tracing over the entire black hole spacetime which thus forms the entanglement wedge of B𝐵B. In the limit keSBHmuch-less-than𝑘superscript𝑒subscript𝑆𝐵𝐻k\ll e^{S_{BH}}, the fully connected replica wormhole geometry (one single disc with n𝑛n asymptotic boundaries) dominates. So we obtain Tr(ρR)=SBHTrsubscript𝜌𝑅subscript𝑆𝐵𝐻{\rm Tr}(\rho_{R})=S_{BH} when n1𝑛1n\rightarrow 1. This implies that the entanglement wedge of R𝑅R is the interior of the horizon which forms the island, while the black hole exterior forms the entanglement wedge of B𝐵B. Figure from Penington:2019kki .

Recently, Bousso and Shahbazi-Moghaddam have revisited the holographic entropy bounds discussed in the introduction and have examined them in the context of islands Bousso:2021sji . In particular, they have been able to find some general conditions for the existence of islands and how holographic entropy bounds should be revised in the context of island rule. They have argued that such arguments could guarantee behavior consistent with unitarity in more general contexts. This also explains absence of islands Manu:2020tty found in the context of AdS-Kasner spacetimes coupled to non-gravitating reservoirs.

The Page curve for other entanglement measures such as entanglement negativity (which is a more suitable measure than the von Neumann entropy in the case of mixed states) of the Hawking radiation leaking to a bath has also been recently reproduced in these two-dimensional setups KumarBasak:2020ams ; KumarBasak:2021rrx developing on prescriptions obtained in Kudler-Flam:2018qjo ; Kusuki:2019zsp ; KumarBasak:2020ams and are consistent with results obtained from random matrix theory Shapourian:2020mkc . The entanglement negativity has also been studied in the context of a generalized version of the PSSY model with a bipartite (non-gravitating) reservoir in Kudler-Flam:2021efr ; Dong:2021oad . Interestingly, saddles which break replica symmetry appear. The Page curve in doubly holographic setups with the (non-gravitating) reservoir subjected to relevant deformations have been studied in Caceres:2021fuw . In this work, it has been shown that the coarse-graining of the reservoir generated by the RG flow leads to an increase in the Page time.

Finally, the emergence of islands should be understood from modular flow which is the fundamental element of explicit entanglement wedge reconstruction as discussed in section 3. See Chen:2019iro for advances made in this direction.

The fact that the replica wormhole saddles of the Euclidean path integrals including the gravitating regions perform some kind of averaging is a generic feature which should hold for higher dimensional setups also. In a full unitary quantum gravity computation in real time in which we should be able to take into account the microstates of the black hole (such as fuzzballs) explicitly, the averaging should emerge dynamically via quantum ergodicity. The late-time self-averaging in many-body dynamics which justifies the Euclidean computation of Rènyi entropies leading to an universal approximation has been discussed in PRXQuantum.2.010344 212121See also Vardhan:2021npf for a related discussion on how such equilibrium approximation of chaotic many-body dynamics makes novel predictions for entanglement in Hawking radiation before Page time.. Nevertheless, it has to be understood from the bulk point of view.222222In fact JT gravity has a full non-perturbative dual description in terms of an ensemble of random matrices Saad:2019lba ; Stanford:2019vob . That such randomness eg random couplings should be described by wormholes has been already discussed by Giddings:1988wv ; Polchinski:1994zs . In an unitary description, such averaging should emerge from ergodicity and then the question is to understand the dual bulk mechanism from explicit microstate models.

It is already remarkable that the replica wormholes produce results for the Page curve which are compatible with unitarity. Nevertheless, many crucial features of the encoding of the black hole interior into the Hawking quanta and other related phenomena such as quantum information mirroring cannot be addressed by such an approach. We will examine very soon if we can construct tractable microstate models to address these fundamental issues.

5.2 State-dependence in microstate reconstruction: Alpha bits and Python’s lunch

The natural question to ask of course is how do we reconstruct (microstate of) a non-evaporating black hole which is not in contact with an auxiliary reservoir.232323By non-evaporating we mean with reflecting asymptotic boundary conditions. Note Hawking radiation is present but the black hole cannot lose mass. This should be formulated in the general context of entanglement wedge reconstruction. This question was originally discussed in the context of the resolution of the AMPS paradox. It was argued by Papadodimas and Raju Papadodimas:2012aq ; Papadodimas:2013jku ; Raju:2020smc that the reconstruction of the black hole interior in the dual boundary theory has to be necessarily state-dependent and this can resolve the AMPS and related information paradoxes. The operators needed for the reconstruction of the interior should themselves be complicated and would be also microstate dependent. Its consistency with the framework of quantum mechanics has been debated, see Harlow:2014yoa for an instance. We will review these issues pertaining to the AMPS paradox in the next section. In this section, we will present developments related to state-dependence of black hole interior reconstruction which has been reformulated in the framework of quantum error correction with crucial inputs from the extremal surfaces in Hayden:2018khn ; Akers:2019wxj ; penington2020entanglement which is manifestly consistent with the basic postulates of quantum mechanics. We will also discuss how this formulation leads to the quantification of the complexity of the reconstruction of the interior Brown:2019rox ; Engelhardt:2021qjs that is necessary for the resolution of the AMPS and related paradoxes (more on this later). Note this discussion relates to the properties of the entanglement wedge and is also applicable to evaporating black holes.

Refer to caption
Figure 23: In presence of a horizon, there are multiple extremal surfaces attached to the boundary A𝐴\partial A of a boundary subregion. As shown above, there are two with areas 𝒜2subscript𝒜2\mathcal{A}_{2} and 𝒜1subscript𝒜1\mathcal{A}_{1}. If the boundary subregion inlcudes more than half of the full boundary, then 𝒜2>𝒜1subscript𝒜2subscript𝒜1\mathcal{A}_{2}>\mathcal{A}_{1}. However, 𝒜1subscript𝒜1\mathcal{A}_{1} is homologous to A¯¯𝐴\overline{A} and 𝒜2subscript𝒜2\mathcal{A}_{2} is homologous to A𝐴A. The homology constraint implies that the entanglement wedge of A𝐴A is a𝑎a, the bulk region bounded by A𝐴A and 𝒜2subscript𝒜2\mathcal{A}_{2} (and that of A¯¯𝐴\overline{A} is similarly a¯¯𝑎\overline{a}) in the thermal state dual to a black hole. The bulk region asuperscript𝑎a^{\prime} bounded by the two minimal surfaces, which includes portion of the exterior of the black hole, cannot be reconstructed at the boundary. As discussed in text, the situation is different in case of a microstate. Figure from Hayden:2018khn .

The crucial point in Hayden:2018khn is that state-dependence in the (approximate) reconstruction of the interior occurs when we consider a boundary sub-region instead of the whole dual system. Furthermore, state dependence actually implies dependence on the code subspace and not dependence on the specific bulk density matrix with support in this code subspace. Geometrically it originates from the dependence of the entanglement wedge on the choice of the code subspace and it is simply determined by the maximally mixed state in this code subspace. To illustrate this form of state-dependence in the reconstruction of the interior, consider a region A𝐴A at the boundary as shown in Fig. 23. There are two extremal RT surfaces with areas 𝒜1subscript𝒜1\mathcal{A}_{1} and 𝒜2subscript𝒜2\mathcal{A}_{2} respectively. Here 𝒜2subscript𝒜2\mathcal{A}_{2} is homologous to A𝐴A and 𝒜1subscript𝒜1\mathcal{A}_{1} to its complement A¯¯𝐴\overline{A}. If A𝐴A is larger than half the boundary, then 𝒜2>𝒜1subscript𝒜2subscript𝒜1\mathcal{A}_{2}>\mathcal{A}_{1}. The bulk is separated into three regions a𝑎a, asuperscript𝑎a^{\prime} and a¯¯𝑎\overline{a} where a(a¯)𝑎¯𝑎a(\overline{a}) is the region bounded by 𝒜2(𝒜1)subscript𝒜2subscript𝒜1\mathcal{A}_{2}(\mathcal{A}_{1}) and A(A¯)𝐴¯𝐴A(\overline{A}), while asuperscript𝑎a^{\prime} is the region between the two minimal surfaces containing the black hole horizon with area 𝒜0subscript𝒜0\mathcal{A}_{0} and portions of the black hole exterior also. Furthermore, 𝒜2𝒜1<𝒜0subscript𝒜2subscript𝒜1subscript𝒜0\mathcal{A}_{2}-\mathcal{A}_{1}<\mathcal{A}_{0} and so we can define

α=𝒜2𝒜1𝒜0𝛼subscript𝒜2subscript𝒜1subscript𝒜0\alpha=\frac{\mathcal{A}_{2}-\mathcal{A}_{1}}{\mathcal{A}_{0}}

so that 0α10𝛼10\leq\alpha\leq 1. The code subspace is

code=aaa¯.subscriptcodetensor-productsubscript𝑎subscriptsuperscript𝑎subscript¯𝑎\mathcal{H}_{\rm code}=\mathcal{H}_{a}\otimes\mathcal{H}_{a^{\prime}}\otimes\mathcal{H}_{\overline{a}}.

In the semi-classical limit G0𝐺0G\rightarrow 0, the dimension of the full code subspace is

eSBH=e𝒜04G.superscript𝑒subscript𝑆BHsuperscript𝑒subscript𝒜04𝐺e^{S_{\rm BH}}=e^{\frac{\mathcal{A}_{0}}{4G}}.

Consider the thermal state dual to the actual black hole geometry. Then the homology constraint implies that the entanglement wedge of A𝐴A is a𝑎a and that of A¯¯𝐴\overline{A} is a¯¯𝑎\overline{a}. Any bulk operator 𝒪asubscript𝒪superscript𝑎\mathcal{O}_{a^{\prime}} localized in asubscriptsuperscript𝑎\mathcal{H}_{a^{\prime}} whether it is in the interior of the black hole or in the exterior cannot be reconstructed either in Asubscript𝐴\mathcal{H}_{A} or in A¯subscript¯𝐴\mathcal{H}_{\overline{A}}. This is not surprising at all. Firstly note that only asubscriptsuperscript𝑎\mathcal{H}_{a^{\prime}} which contains the horizon region will have dimension 𝒪(e1/G)𝒪superscript𝑒1𝐺\mathcal{O}(e^{1/G}) whereas asubscript𝑎\mathcal{H}_{a} and a¯subscript¯𝑎\mathcal{H}_{\overline{a}} will have dimension 𝒪(1)𝒪1\mathcal{O}(1) in the limit G0𝐺0G\rightarrow 0. This implies that none of the microstates can be resolved as should be the case in the thermal ensemble. Only in the case of a microstate, we should be able to access (reconstruct) states in asubscriptsuperscript𝑎\mathcal{H}_{a^{\prime}} at the boundary.

To be precise, we assume that for any typical microstate we can consider the same semi-classical geometry up to where the surfaces 𝒜1subscript𝒜1\mathcal{A}_{1} and 𝒜2subscript𝒜2\mathcal{A}_{2} are located (indeed valid for fuzzballs which differ from the black hole significantly only at the scale of the horizon Mathur:2005zp ). This assumption is however not crucial for what follows – it is enough if the location of 𝒜1subscript𝒜1\mathcal{A}_{1} and 𝒜2subscript𝒜2\mathcal{A}_{2} have sub-leading state dependence. Crucially, in a microstate (with a smooth geometry) there will be no homology constraint. For reasons which will be clear later, we should consider a mixed state in the bulk. In order to do this, let us entangle the black hole with a reference system R𝑅R (note this does not imply that they are in physical contact or R𝑅R is acting as a reservoir of Hawking quanta). Let |ψket𝜓|\psi\rangle be a pure state in codeRtensor-productsubscriptcodesubscript𝑅\mathcal{H}_{\rm code}\otimes\mathcal{H}_{R}. The question is then can we decode an operator 𝒪asubscript𝒪superscript𝑎\mathcal{O}_{a^{\prime}}, localized in asubscriptsuperscript𝑎\mathcal{H}_{a^{\prime}}, in A𝐴A. It will be possible if the entanglement wedge of AR𝐴𝑅A\cup R is aaR𝑎superscript𝑎𝑅a\cup a^{\prime}\cup R. This would be so if Sgen(𝒜2)>Sgen(𝒜1)subscript𝑆𝑔𝑒𝑛subscript𝒜2subscript𝑆𝑔𝑒𝑛subscript𝒜1S_{gen}(\mathcal{A}_{2})>S_{gen}(\mathcal{A}_{1}) implying that the quantum extremal surface corresponding to A𝐴A should be 𝒜1subscript𝒜1\mathcal{A}_{1} instead of 𝒜2subscript𝒜2\mathcal{A}_{2} for all states |ψcodeRket𝜓tensor-productsubscriptcodesubscript𝑅|\psi\rangle\in\mathcal{H}_{\rm code}\otimes\mathcal{H}_{R}. Therefore, we need (using S(aa)ψ=S(a¯R)ψ𝑆subscript𝑎superscript𝑎𝜓𝑆subscript¯𝑎𝑅𝜓S(aa^{\prime})_{\psi}=S(\overline{a}R)_{\psi} for a pure state |ψket𝜓|\psi\rangle):

S(a)ψ+𝒜24G>S(aa)ψ+𝒜14G=S(a¯R)ψ+𝒜14G.𝑆subscript𝑎𝜓subscript𝒜24𝐺𝑆subscript𝑎superscript𝑎𝜓subscript𝒜14𝐺𝑆subscript¯𝑎𝑅𝜓subscript𝒜14𝐺S(a)_{\psi}+\frac{\mathcal{A}_{2}}{4G}>S(aa^{\prime})_{\psi}+\frac{\mathcal{A}_{1}}{4G}=S(\overline{a}R)_{\psi}+\frac{\mathcal{A}_{1}}{4G}. (98)

As discussed above, in the limit G0𝐺0G\rightarrow 0, S(a)ψ𝑆subscript𝑎𝜓S(a)_{\psi} is 𝒪(1)𝒪1\mathcal{O}(1). Also the triangle inequality implies

|S(R)ψS(a¯R)ψ|S(a¯)ψ=𝒪(1).𝑆subscript𝑅𝜓𝑆subscript¯𝑎𝑅𝜓𝑆subscript¯𝑎𝜓𝒪1|S(R)_{\psi}-S(\overline{a}R)_{\psi}|\leq S(\overline{a})_{\psi}=\mathcal{O}(1).

So (98) reduces to the condition that if

4GS(R)ψ<𝒜2𝒜1,4𝐺𝑆subscript𝑅𝜓subscript𝒜2subscript𝒜14GS(R)_{\psi}<\mathcal{A}_{2}-\mathcal{A}_{1}, (99)

then the entanglement wedge of A𝐴A is aa𝑎superscript𝑎a\cup a^{\prime}. This requires that |ψket𝜓|\psi\rangle belongs to a code subspace 𝒮codesubscript𝒮subscriptcode\mathcal{H}_{\mathcal{S}}\subseteq\mathcal{H}_{\rm code} of dimension dS=dRsubscript𝑑𝑆subscript𝑑𝑅d_{S}=d_{R} such that the above inequality is satisfied for any state in this subspace. The reconstruction criterion then arises from the maximally mixed state in the subspace for which S(R)ψ=logdS𝑆subscript𝑅𝜓subscript𝑑𝑆S(R)_{\psi}=\log d_{S} implying that

dS<e𝒜2𝒜14G=eαSBH.subscript𝑑𝑆superscript𝑒subscript𝒜2subscript𝒜14𝐺superscript𝑒𝛼subscript𝑆BHd_{S}<e^{\frac{\mathcal{A}_{2}-\mathcal{A}_{1}}{4G}}=e^{\alpha S_{\rm BH}}. (100)

Irrespective of whichever state we choose in this subspace 𝒮subscript𝒮\mathcal{H}_{\mathcal{S}}, any bulk operator 𝒪asubscript𝒪superscript𝑎\mathcal{O}_{a^{\prime}} localized in asubscriptsuperscript𝑎\mathcal{H}_{a^{\prime}} can be reconstructed via the same operator 𝒪Asubscript𝒪𝐴\mathcal{O}_{A} in Asubscript𝐴\mathcal{H}_{A}. Nevertheless, this operator 𝒪Asubscript𝒪𝐴\mathcal{O}_{A} will depend on the choice of the code subspace – outside this subspace there exists states (eg. the fully entangled state of the black hole interior and R𝑅R) the action of 𝒪asubscript𝒪superscript𝑎\mathcal{O}_{a^{\prime}} on which cannot be simulated by 𝒪Asubscript𝒪𝐴\mathcal{O}_{A} since the entanglement wedge of A𝐴A does not contain asuperscript𝑎a^{\prime} for such states. The state-dependence of 𝒪Asubscript𝒪𝐴\mathcal{O}_{A} is via the choice of this appropriate code subspace alone. As the size of A𝐴A grows, clearly α1𝛼1\alpha\rightarrow 1. Therefore, the code subspace involves all typical states in the full Hilbert space and there is no state dependence in the reconstructed operators 𝒪Asubscript𝒪𝐴\mathcal{O}_{A} which have the same action as 𝒪asubscript𝒪superscript𝑎\mathcal{O}_{a^{\prime}} acting in the interior of the black hole.

The quantum error correcting protocol which reproduces the above desired behavior is the universal subsystem recovery channel first considered in Hayden:2017xed . Suppose there is a quantum channel that applies a Haar-random unitary U𝑈U to n𝑛n qubits, throws away a fraction of them and transmits the rest. Let the input Hilbert space be insubscriptin\mathcal{H}_{\rm in}. If one retains a fraction (1+α)/21𝛼2(1+\alpha)/2 of the qubits, then one will be able to decode (recover) any subspace in the input Hilbert space which has αn𝛼𝑛\alpha n qubits (this subspace therefore has dimension 2αnsuperscript2𝛼𝑛2^{\alpha n}.) The universal subspace recovery map however can only approximately reverse the channel. Sometimes, there has to be an error ϵ>eηnitalic-ϵsuperscript𝑒𝜂𝑛\epsilon>e^{-\eta n} with η>0𝜂0\eta>0. In the context of reconstruction of microstates this implies error

ϵ>eη/Gitalic-ϵsuperscript𝑒𝜂𝐺\epsilon>e^{-\eta/G}

which should be typically non-perturbatively small in G𝐺G. The explicit universal recovery map is closely related to the twirled Petz map Hayden:2018khn which has tantalizing connections with the modular Hamiltonian as discussed before. The bulk region asuperscript𝑎a^{\prime} between the minimal surfaces contains the α𝛼\alpha-bits in the language of universal subspace error correction.

Such issues in bulk reconstruction arise not only in the case of black holes but also when such competing extremal surfaces enclose bulk matter with high entropy – we refer the reader to Akers:2019wxj ; Akers:2020pmf for extensive discussions. In particular Akers:2020pmf discusses how the quantum extremal surface prescriptions should be refined in such generic circumstances with information-theoretic interpretations.

Intuitively it should be hard to reconstruct operators lying in asuperscript𝑎a^{\prime} between the two minimal surfaces in Fig 23 from A𝐴A even within the code subspace essentially due to state-dependence. This is similar to the Hayden-Harlow conjecture of exponential complexity of reconstruction of operators in the interior of the black hole restricted to Hawking radiation to be discussed in the following subsection. A remarkable geometric way of capturing this complexity was proposed in Brown:2019rox based on the generic existence of non-minimal extremal surfaces forming a Python-lunch geometry (sandwiched between minimal extremal surfaces or behind one of them) and interpreting this geometry in terms of tensor network constructions. This proposal further develops earlier proposals for the holographic dictionary for complexity Susskind:2014rva ; Brown:2015bva ; Brown:2015lvg . For a recent review on quantum complexity and its holographic description see Shira_review .

The mechanism for nucleation for such Python lunches was proposed in Engelhardt:2021qjs . Essentially the argument in the case of the black hole setup of Fig 23 is that we should first consider the appropriate code subspace where the interior can be decoded in A𝐴A. In the maximally mixed state in this code-subspace the Hawking pairs will experience disentanglement and the bulk entropy gradients will be larger than the Hartle-Hawking state. Using results from Wall:2012uf ; Marolf:2019bgj ; Akers:2019lzs , then it has been argued that due to blueshift of the entropy gradients when extrapolated in the past there would be nucleations of non-minimal (highly non-classical) extremal surfaces (since it can compete with the leading order 1/G1𝐺1/G term) behind the horizon ensuring that they do exist also at late time (this is somewhat similar to the existence of the transition in the extremal surfaces we have seen before in the context of reproduction of the Page curve). Even if one considers the full boundary, one needs to consider a code subspace for discussing the interior outgoing modes and then there will be such a bulge surface γbulgesubscript𝛾bulge\gamma_{\rm bulge} behind the outermost extremal surface γaptzsubscript𝛾aptz\gamma_{\rm aptz} (the appetizer of the Python’s lunch) – see Fig. 24. The complexity of the decoding (defined below) within this code subspace will then be half the difference in the generalized entropies of γbulgesubscript𝛾bulge\gamma_{\rm bulge} and γaptzsubscript𝛾aptz\gamma_{\rm aptz}.

Refer to caption
Figure 24: In order to consider excited modes in the black hole interior, one first needs to restrict to a code subspace where the geometry will have a Python’s lunch – a non-minimal extremal surface γbulgesubscript𝛾bulge\gamma_{\rm bulge} behind the outermost minimal extremal surface γaptzsubscript𝛾aptz\gamma_{\rm aptz}. The throat refers to the asymptotic region. There can be other locally minimal surfaces behind γbulgesubscript𝛾bulge\gamma_{\rm bulge} (not shown in figure). Figure from Engelhardt:2021qjs .

For the moment, let us assume that such a Python lunch geometry is generic when one considers appropriate code subspaces where interior modes are excited and can be defined in terms of the macroscopic geometry corresponding to the maximally mixed state in this code subspace. In this case, one can justify the quantification of the complexity of decoding the interior from tensor network models. Such models were also used to derive the holographic complexity conjectures Susskind:2014rva ; Brown:2015bva ; Brown:2015lvg which need to be modified in the presence of the Python lunches. For an illustration consider such a wormhole with a Python lunch in the two-sided thermofield double geometry represented in the form of a tensor network in Fig. 25. We discuss the circuit complexity of decoding following Engelhardt:2021qjs .

Refer to caption
Figure 25: A tensor network dual to a Python’s lunch geometry such that the map from the input state in the left Hilbert space and the black dots representing bulk legs to the right Hilbert space is an approximate isometry. It has two locally minimal cuts γaptzsubscript𝛾aptz\gamma_{\rm aptz} and γminsubscript𝛾min\gamma_{\rm min} with γaptz>γminsubscript𝛾aptzsubscript𝛾min\gamma_{\rm aptz}>\gamma_{\rm min}. Furthermore, there is a maximal cut γbulgesubscript𝛾bulge\gamma_{\rm bulge} between these two cuts. The triangles denote isometries whereas the squares have one or more of their legs projected to |0ket0|0\rangle state. Figure from Engelhardt:2021qjs .

The tensor network represents the map from the bulk represented as black dots in Fig 25 to the tensor product of the left and right CFT Hilbert spaces. Triangles represent isometries and the squares involve postselection – an isometry where one (or more) of the legs is projected to |0ket0|0\rangle. In a generic tensor network of such type, there will be two locally minimal cuts γminsubscript𝛾min\gamma_{\rm min} (global minimum) and γaptzsubscript𝛾aptz\gamma_{\rm aptz} (representing the locally minimal extremal surfaces) and a γbulgesubscript𝛾bulge\gamma_{\rm bulge} locally maximal cut in the middle. When we view the figure from the left to right it represents an approximate isometry from the tensor product of bulk legs and left CFT Hilbert space to the right CFT Hilbert space. For this to be the case, we need to assume that the bonds cut in γaptzsubscript𝛾aptz\gamma_{\rm aptz} have larger dimension than those in the γminsubscript𝛾min\gamma_{\rm min} cut plus the dimensions of the bulk legs (denoted as black dots) between these two cuts. However, the problem is in the middle because the map from the γbulgesubscript𝛾bulge\gamma_{\rm bulge} cut plus the bulk legs in between the γbulgesubscript𝛾bulge\gamma_{\rm bulge} and the γaptzsubscript𝛾aptz\gamma_{\rm aptz} cut to the γaptzsubscript𝛾aptz\gamma_{\rm aptz} cut is not an isometry as the Hilbert space dimension of the latter is smaller. To make it work in practice we need to somehow implement the postselection (which is not unitary as it involves projection) in a different way via unitary circuits. This can be done via Grover search algorithm in the manner discussed first in Yoshida:2017non – we need to do sequential unitary transformations such that we bring the qubits that are supposed to be postselected already in |0ket0|0\rangle state. The complexity of the decoding of the Python lunch is essentially that of this Grover search algorithm. By general arguments this complexity is

C=O(C~2m/2)𝐶𝑂~𝐶superscript2𝑚2C=O(\tilde{C}2^{m/2})

where m𝑚m is the number of qubits which should be postselected (equal to the difference between the bulge and appetizer cuts) and C~~𝐶\tilde{C} is related to the overall size of the network.242424For a discussion on complexity utilizing the Petz map in the context of a subregion see Zhao:2020wgp .

Now in the gravitational analogue γaptzsubscript𝛾aptz\gamma_{\rm aptz} and γminsubscript𝛾min\gamma_{\rm min} are surfaces with locally minimal generalized entropies. By our earlier discussion the code subspace that can be reconstructed in the right Hilbert space has dimension

Sgen(γaptz)Sgen(γmin)subscript𝑆𝑔𝑒𝑛subscript𝛾aptzsubscript𝑆𝑔𝑒𝑛subscript𝛾minS_{gen}(\gamma_{\rm aptz})-S_{gen}(\gamma_{\rm min})

with generalized entropies defined by considering the maximally mixed state in the code subspace. The condition Sgen(γaptz)>Sgen(γmin)subscript𝑆𝑔𝑒𝑛subscript𝛾aptzsubscript𝑆𝑔𝑒𝑛subscript𝛾minS_{gen}(\gamma_{\rm aptz})>S_{gen}(\gamma_{\rm min}) is the reconstructibility criterion analogous to that for the tensor network to be an isometry to the right Hilbert space. Furthermore, γbulgesubscript𝛾bulge\gamma_{\rm bulge} has locally maximal generalized entropy and for the Python lunch geometry Sgen(γbulge)>Sgen(γaptz)subscript𝑆𝑔𝑒𝑛subscript𝛾bulgesubscript𝑆𝑔𝑒𝑛subscript𝛾aptzS_{gen}(\gamma_{\rm bulge})>S_{gen}(\gamma_{\rm aptz}). Then the tensor network analogy suggests complexity of decoding should be

C=𝒪(C~exp12(Sgen(γbulge)Sgen(γaptz))).𝐶𝒪~𝐶12subscript𝑆𝑔𝑒𝑛subscript𝛾bulgesubscript𝑆𝑔𝑒𝑛subscript𝛾aptzC=\mathcal{O}(\tilde{C}\exp\frac{1}{2}\left(S_{gen}(\gamma_{\rm bulge})-S_{gen}(\gamma_{\rm aptz}))\right).

The exponential dependence in G1superscript𝐺1G^{-1} comes crucially only from the difference of the generalized entropies while the dependence on G1superscript𝐺1G^{-1} in C~~𝐶\tilde{C} (which is of same order in G1superscript𝐺1G^{-1} as the generalized entropies) is only linear. So the former accounts for exponential complexity. A similar discussion can be repeated for a single sided microstate geometry shown in Fig 24 also.

The understanding of the mechanism for generation of exponential complexity already alerts us towards the need for more explicit microstate geometries that can describe appropriate code subspaces. We will further motivate microstate models in the following subsection.

Interestingly a converse of the Python’s lunch conjecture has been discussed in Engelhardt:2021mue . It has been claimed that bulk operators between the boundary and the outermost extremal surface should have a simple reconstruction at the boundary in the sense that they can be recovered efficiently from a dual coarse-grained state with an effective local modular Hamiltonian. See also Levine:2020upy for a related discussion. Recent discussions of the reconstruction of the experience of a bulk observer in the dual conformal field theory can be found in Jafferis:2020ora via the use of modular flow and through emergent properties of von-Neumann algebras in Leutheusser:2021frk .

5.3 Decoding the interior in real time

5.3.1 Quantum information mirroring and the resolution of the AMPS paradox via complexity

Quantum information theory has been applied to understand how the black hole acts as a quantum channel with a motivation to resolve the AMPS paradox. Such analyses predict very non-trivial features of how the interior gets encoded into the outgoing Hawking quanta in real time.

An extension of Page’s argument implies that old black holes (past its Page time) would act as quantum information mirrors following a thought experiment and its analysis due to Hayden and Preskill Hayden_2007 as illustrated in Fig. 26. A typical state of an old black hole would have Hawking quanta in the exterior (E𝐸E) maximally entangled with the interior modes (B𝐵B). If the black hole is a fast scrambler, then qubits (D𝐷D) thrown into it would be maximally scrambled with B𝐵B (one could understand this as an action of a random unitary operator U𝑈U252525Hayden and Preskill showed that this unitary operator need not have exponentially large number of gates for a Page-like argument to work. It suffices to pick the unitary randomly from a unitary two-design, which can be achieved by a quantum circuit of depth 𝒪(logn)𝒪𝑛\mathcal{O}(\log n) where n=SBH𝑛subscript𝑆𝐵𝐻n=S_{BH} is the total number of qubits at the horizon with S𝑆S being the black hole entropy. If each step in the circuit takes Plank time, then it should be redshifted to rssubscript𝑟𝑠r_{s} in the time of the asymptotic observer. It then reproduces the scrambling time as the circuit time which is rslogSBHsubscript𝑟𝑠subscript𝑆𝐵𝐻r_{s}\log S_{BH}. This argument was refined in Sekino:2008he ) resulting in a remnant Bsuperscript𝐵B^{\prime} and newly radiated Hawking quanta R𝑅R. For a formal statement, we consider that D𝐷D is maximally entangled with a reference system S𝑆S, so that post-scrambling an extension of Page’s arguments would imply that S𝑆S and Bsuperscript𝐵B^{\prime} have no mutual information (entanglement). However, the information in D𝐷D would be transferred to the combined RE𝑅𝐸R\cup E system of Hawking quanta implying information mirroring which would essentially happen at scrambling time. The latter has been computed based on growth of thermal commutators by Shenker and Stanford Shenker:2013pqa . In the context of black holes, it is essentially the time (from the point of view of an asymptotic observer) it takes a light ray to reach Planck distance close to the stretched horizon, which is rslogSBHabsentsubscript𝑟𝑠subscript𝑆𝐵𝐻\approx r_{s}\log S_{BH}, with SBHsubscript𝑆𝐵𝐻S_{BH} being the entropy of the black hole. For a detailed review of the Hayden-Preskill protocol and progress on experimental realization of quantum simulators which achieves the scrambling needed to realize it see Lata_review .

Refer to caption
Figure 26: An illustration of the Hayden-Preskill thought experiment which argues for quantum information mirroring in old black holes – figure taken from Harlow:2014yka . A quantum diary D𝐷D entangled with a reference state S𝑆S is thrown into an old black hole B𝐵B which is entangled with its radiation E𝐸E. The diary D𝐷D and B𝐵B then gets scrambled by the unitary evolution operator U𝑈U after which we obtain the remaining black hole Bsuperscript𝐵B^{\prime} and some more radiation R𝑅R. If R𝑅R has only a few more qubits than D𝐷D, then we can show that mutual information between S𝑆S and Bsuperscript𝐵B^{\prime} should be negligible given that U𝑈U is random. The information in D𝐷D is then in the ER𝐸𝑅E\cup R system after the scrambling time.

The natural question that arises is how easy it would be to decode the qubits in D𝐷D from ER𝐸𝑅E\cup R. Before considering this question we need to encounter first the most pragmatic operational way to resolve the AMPS paradox that leads to an additional feature of the encoding of the interior into the Hawking quanta. In order to present this, it is useful to restate the AMPS paradox. Once again consider the old black hole for which the interior modes B𝐵B should be maximally entangled with E𝐸E, the already radiated quanta. Unitarity of the time-evolution would imply that the newly emitted quanta R𝑅R that decouples from the black hole can be purified by a factor ERsubscript𝐸𝑅E_{R} of E𝐸E so that ρRERsubscript𝜌𝑅subscript𝐸𝑅\rho_{RE_{R}} is a pure state. However, R𝑅R must also be maximally entangled with (a factor of) B𝐵B if semiclassical EFT holds at the horizon. A simple way to resolve this paradox is to invoke the philosophy of black hole complementarity PhysRevD.48.3743 ; PhysRevD.50.2700 which postulates that the violation cannot be observed operationally via a quantum complexity conjecture due to Harlow and Hayden Harlow:2013tf (see Aaronson:2016vto for a nice discussion). Essentially this argument states that in order to distill the factor ERsubscript𝐸𝑅E_{R} from E𝐸E needed to purify R𝑅R, it would take time which is exponential in the entropy of the black hole (at the time when R𝑅R was emitted). This conjecture can be made more precise by evoking pseudorandom encoding of the interior into the outgoing Hawking quanta Kim:2020cds (in microstate models we will be able to relate inherent pseudorandom/chaotic dynamics with Python lunches that can macroscopically amplify small excitations). Since the black hole evaporates in time that is polynomial in its entropy, the violation of monogamy of entanglement cannot be demonstrated operationally.

This discussion raises a few fundamental questions. Firstly, what exactly does an operational resolution mean? Does it mean we need to modify the framework of quantum mechanics actually to describe black holes although we cannot operationally test violation of its postulates? Or does it mean that somehow these Hilbert spaces, especially B𝐵B (the interior) and E𝐸E (the pre-Page time Hawking quanta) are actually not separable but only in an operational sense? Then in this operational framework would the AMPS paradox be resolved by complexity? The Euclidean replica wormhole saddles which connect E𝐸E and B𝐵B would support the latter point of view. However, could we actually understand how to make sense of such an operational framework in real time? The latter is challenging as we would also need to validate the usual semi-classical picture of the black hole (at least from the point of view of measurements of the EFT observables). In fact, the Page curve can be computed using only the semi-classical geometry as discussed before and therefore it should be valid for a large class of measurements.

The second class of issues are related with the encoding of the information in Hawking radiation. If rapid mirroring of the information thrown in after Page time could happen together with the complex encoding of the Hawking interior, then which of these two possibilities could be true: (a) one can decode the qubits D𝐷D thrown into the black hole from the Hawking quanta immediately after scrambling time without explicit knowledge of the interior encoded in E𝐸E (decoding the latter would take time exponential in the entropy of the black hole), or (b) we can only decode it after we actually know the interior after an enormously long time. The discussion on islands in the form of entanglement wedge of R𝑅R would actually prefer the first possibility. In the setups discussed in the previous subsection, the newly emitted Hawking quanta would be in the entanglement wedge of the holographic system B𝐵B and not yet in the bath region. The physical separation between these regions would indicate that the mirrored information could be readily decoded soon after the new Hawking quanta emerges in the entanglement wedge of B𝐵B. However, one still needs to demonstrate that the decoding is possible without the knowledge of the island (the black hole interior) that is encoded in this new radiation since the knowledge of how the island has been modified by the infalling bits should eventually leak into the bath. Therefore, one needs to identify such features of the encoding of the infalling qubits which co-exist with the complex encoding of the interior, and explain their physical origins. One could formulate these questions also more generally without the setups of the previous subsection by advocating an emergent infrared holographic theory that describes the near-horizon geometry of the black hole and which could play the role of B𝐵B.

In the next subsection, we will review a microstate model that could be promising for finding answers to such questions.

5.3.2 Microstate dynamics: Towards understanding quantum black holes in real time

It is useful to study tractable models of black hole microstate dynamics to gain insights into how the features of quantum information mirroring and complex encoding of interior appear in the Hawking radiation, and also for understanding how the principle of black hole complementarity and the averaging implied in replica wormholes could emerge operationally. The fuzzball program Mathur:2005zp , if developed to its full potential, would be able to reveal these mysteries. However, at present, it will be very difficult to study the quantum dynamics of fuzzballs in a sufficiently detailed way. The same could be said about large N𝑁N BFSS matrix models Banks:1996vh , etc. We will argue that certain simplified models could be promising for understanding many (if not all) aspects of these issues. This class of models described in PhysRevD.102.086008 essentially simplify some of the otherwise untractable aspects of quantum gravity in a suitable way, and can be studied in the same spirit as the setups described earlier in this review. One key aspect of this class of models is to give prominence to hair degrees of freedom on the horizon which are essentially (approximately) conserved (non-)gravitational charges whose role in the information paradox have been emphasised in PhysRevLett.116.231301 ; Strominger:2017aeh . The crucial and novel aspect of the hair that will be of interest to us is how it can enable an operational definition of separability of the interior and exterior Hilbert spaces while resolving AMPS type paradoxes that threaten black hole complementarity, and enable mechanisms for the complex encoding of interior, and quantum information mirroring where decoding could be possible without the knowledge of the encoding of the interior.

A class of models: The setup of these microstate models can be motivated from the fragmentation instability Maldacena:1998uz of the near-horizon geometry of the near-extremal black hole mediated by the Brill instantons of semi-classical gravity. These imply the fragmentation of the near-horizon geometry into several two-dimensional throats of the type AdS2×X𝐴𝑑subscript𝑆2𝑋AdS_{2}\times X, where X𝑋X is a compact space that has the topology of the horizon and AdS2𝐴𝑑subscript𝑆2AdS_{2} is the two-dimensional anti-de Sitter space. At the boundaries of the instanton moduli space where the centers of some of the throats come within Planckian distance proximity to each other, there is a proliferation of soft modes. However, quantum gravity effects also become large, so the semi-classical computation becomes intractable. Heuristically, we can assume that these fragmented throats crystallise into a stable configuration forming a lattice of AdS2𝐴𝑑subscript𝑆2AdS_{2} spaces as shown in Fig. 27. These throats should interact with each other via mobile hair which are the (non-)gravitational charges of the original unfragmented part of the geometry and other soft modes that live here. The lattice is simpy a discretization of the compact X𝑋X space that has the same topology of the horizon. Additionally, we can allow the AdS2𝐴𝑑subscript𝑆2AdS_{2} spaces to join in their interiors and form complex networks. Here, we will restrict to the simplest version where the AdS2𝐴𝑑subscript𝑆2AdS_{2} throats do not join with each other and are infinitely extended in the ingoing Eddinton-Finkelstein radial coordinate.

Refer to caption
Figure 27: An illustration of the microstate models comprising of a lattice of two-dimensional AdS2𝐴𝑑subscript𝑆2AdS_{2} throats coupling to each other via mobile hair carrying gravitational charges. The lattice is a discretization of the horizon, a compact space. The AdS2𝐴𝑑subscript𝑆2AdS_{2} throats may join in the interior forming complex networks. Hawking evaporation occurs due to appropriate asymptotic boundary conditions at the throats. This figure is from PhysRevD.102.086008 .

For further tractabilibity and simplification, we consider that each AdS2𝐴𝑑subscript𝑆2AdS_{2} throat in the lattice is described a by semiclassical JT-gravity theory with conformal bulk matter. The semi-classical picture is thus valid for the observer infalling at any throat, however non-local measurements will see struture at the horizon. We will soon discuss how the energy absorbing and relaxation dynamics of the semi-classical black hole emerges from this microstate model. We begin our discussion by first arresting Hawking radiation by enforcing the usual reflecting boundary conditions in each AdS2𝐴𝑑subscript𝑆2AdS_{2} throat so that we can study the intrinsic properties of the microstates first. Note that we retain only a coarse-grained (effective infrared) description of each throat which is dual to a quantum dot in the Planckian lattice. Regardless, we should enforce total energy conservation of this combined lattice of quantum dots and mobile hair system in absence of Hawking radiation. Although, we do modify the effective description of horizon physics, usual statistical arguments would imply that the results of typical measurements would be almost the same as in a microcanonical ensemble which we will describe below. For simplicity, we will consider the horizon to be S1superscript𝑆1S^{1}, so the lattice would be a ring (chain) with periodic boundary conditions.

The state of the i𝑖i-th throat (dual to the i𝑖i-th quantum dot) can then be described by ti(u)subscript𝑡𝑖𝑢t_{i}(u) where u𝑢u is the time of the asymptotic observer. Essentially ti(u)subscript𝑡𝑖𝑢t_{i}(u) is the time-reparametrization mode of the i𝑖i-th throat which determines the correlation functions in the localized state in terms of the vacuum correlation functions as in the Sachdev-Ye-Kitaev model Sachdev_1993 ; Kitaev_2015 which could be the (UV complete) theory governing the quantum dots. In other words, ti(u)subscript𝑡𝑖𝑢t_{i}(u) is the time determining the state of the i𝑖i-th throat as a function of a common vacuum state time which is identified with the time of the vacuum observer. For ease of simulations, it is useful to define τi(u)subscript𝜏𝑖𝑢\tau_{i}(u) via

ti(u)=tanh(τi(u)2)subscript𝑡𝑖𝑢subscript𝜏𝑖𝑢2t_{i}(u)=\tanh\left(\frac{\tau_{i}(u)}{2}\right) (101)

where τi(u)subscript𝜏𝑖𝑢\tau_{i}(u) maps to the common time of a thermal state (instead of the vacuum state) with β=2π𝛽2𝜋\beta=2\pi. In absence of coupling between throats and sources for the bulk matter, each individual throat should have conserved SL(2,R)𝑆𝐿2𝑅SL(2,R) charges which are:

𝒬i0superscriptsubscript𝒬𝑖0\displaystyle\mathcal{Q}_{i}^{0} =\displaystyle= τi′′′τi2τi′′2τi3τi,superscriptsubscript𝜏𝑖′′′superscriptsubscript𝜏𝑖2superscriptsuperscriptsubscript𝜏𝑖′′2superscriptsuperscriptsubscript𝜏𝑖3superscriptsubscript𝜏𝑖\displaystyle\frac{\tau_{i}^{\prime\prime\prime}}{\tau_{i}^{\prime 2}}-\frac{{\tau_{i}^{\prime\prime}}^{2}}{{\tau_{i}^{\prime}}^{3}}-\tau_{i}^{\prime},
𝒬i+superscriptsubscript𝒬𝑖\displaystyle\mathcal{Q}_{i}^{+} =\displaystyle= eτi(τi′′′τi2τi′′2τi3τi′′τi),superscript𝑒subscript𝜏𝑖superscriptsubscript𝜏𝑖′′′superscriptsubscript𝜏𝑖2superscriptsuperscriptsubscript𝜏𝑖′′2superscriptsuperscriptsubscript𝜏𝑖3superscriptsubscript𝜏𝑖′′superscriptsubscript𝜏𝑖\displaystyle e^{\tau_{i}}\left(\frac{\tau_{i}^{\prime\prime\prime}}{\tau_{i}^{\prime 2}}-\frac{{\tau_{i}^{\prime\prime}}^{2}}{{\tau_{i}^{\prime}}^{3}}-\frac{\tau_{i}^{\prime\prime}}{\tau_{i}^{\prime}}\right),
𝒬isuperscriptsubscript𝒬𝑖\displaystyle\mathcal{Q}_{i}^{-} =\displaystyle= eτi(τi′′′τi2τi′′2τi3+τi′′τi).superscript𝑒subscript𝜏𝑖superscriptsubscript𝜏𝑖′′′superscriptsubscript𝜏𝑖2superscriptsuperscriptsubscript𝜏𝑖′′2superscriptsuperscriptsubscript𝜏𝑖3superscriptsubscript𝜏𝑖′′superscriptsubscript𝜏𝑖\displaystyle e^{-\tau_{i}}\left(\frac{\tau_{i}^{\prime\prime\prime}}{\tau_{i}^{\prime 2}}-\frac{{\tau_{i}^{\prime\prime}}^{2}}{{\tau_{i}^{\prime}}^{3}}+\frac{\tau_{i}^{\prime\prime}}{\tau_{i}^{\prime}}\right). (102)

We denote these collectively as 𝒬isubscript𝒬𝑖\vec{\mathcal{Q}}_{i}. The Casimir of these SL(2,R)𝑆𝐿2𝑅SL(2,R) charges is the Arnowitt-Deser-Misner (ADM) mass Misubscript𝑀𝑖M_{i} of the AdS2𝐴𝑑subscript𝑆2AdS_{2} throat, i.e.

Mi=𝒬i02𝒬i+𝒬i=2Sch(τi(u),u)+τi2,subscript𝑀𝑖superscriptsuperscriptsubscript𝒬𝑖02superscriptsubscript𝒬𝑖superscriptsubscript𝒬𝑖2Schsubscript𝜏𝑖𝑢𝑢superscriptsuperscriptsubscript𝜏𝑖2M_{i}={\mathcal{Q}_{i}^{0}}^{2}-\mathcal{Q}_{i}^{+}\mathcal{Q}_{i}^{-}=-2\,{\rm Sch}(\tau_{i}(u),u)+{\tau_{i}^{\prime}}^{2}, (103)

where SchSch\rm{Sch} denotes the Schwarzian derivative

Sch(f(u),u)=f′′′f32f′′2f2.Sch𝑓𝑢𝑢superscript𝑓′′′superscript𝑓32superscript𝑓′′2superscript𝑓2{\rm Sch}(f(u),u)=\frac{f^{\prime\prime\prime}}{f^{\prime}}-\frac{3}{2}\frac{f^{\prime\prime 2}}{f^{\prime 2}}. (104)

On top of these lattice SL(2,R)𝑆𝐿2𝑅SL(2,R) charges, we need to consider additionally mobile hair charges which we can take to be SL(2,R)𝑆𝐿2𝑅SL(2,R) charges too representing the gravitational charges of the unfragmented geometry. These we denote as 𝖰isubscript𝖰𝑖\vec{\mathsf{Q}}_{i} which follow discretized Klein-Gordon type equation when decoupled from the lattice charges. We denote AB𝐴𝐵\vec{A}\cdot\vec{B} as the SL(2,R)𝑆𝐿2𝑅SL(2,R) invariant dot product of two SL(2,R)𝑆𝐿2𝑅SL(2,R) vectors A𝐴\vec{A} and B𝐵\vec{B}.262626Explicitly, AB=A0B012(A+B+AB+).𝐴𝐵superscript𝐴0superscript𝐵012superscript𝐴superscript𝐵superscript𝐴superscript𝐵\vec{A}\cdot\vec{B}=A^{0}B^{0}-\frac{1}{2}(A^{+}B^{-}+A^{-}B^{+}). The simplest equations of motion of this system of gravitational lattice charges 𝒬i(u)subscript𝒬𝑖𝑢\vec{\mathcal{Q}}_{i}(u) and mobile hair 𝖰i(u)subscript𝖰𝑖𝑢\vec{\mathsf{Q}}_{i}(u), which reproduce desired phenomenological properties of a classical black hole, take the form:

Misuperscriptsubscript𝑀𝑖\displaystyle M_{i}^{\prime} =\displaystyle= λ(𝒬i1+𝒬i+12𝒬i)𝖰i,𝜆subscript𝒬𝑖1subscript𝒬𝑖12subscript𝒬𝑖superscriptsubscript𝖰𝑖\displaystyle-\lambda(\vec{\mathcal{Q}}_{i-1}+\vec{\mathcal{Q}}_{i+1}-2\vec{\mathcal{Q}}_{i})\cdot\vec{\mathsf{Q}}_{i}^{\prime},
𝖰i′′superscriptsubscript𝖰𝑖′′\displaystyle\vec{\mathsf{Q}}_{i}^{\prime\prime} =\displaystyle= 1σ2(𝖰i1+𝖰i+12𝖰i)1superscript𝜎2subscript𝖰𝑖1subscript𝖰𝑖12subscript𝖰𝑖\displaystyle\frac{1}{\sigma^{2}}(\vec{\mathsf{Q}}_{i-1}+\vec{\mathsf{Q}}_{i+1}-2\vec{\mathsf{Q}}_{i}) (105)
+1λ2(𝒬i1+𝒬i+12𝒬i).1superscript𝜆2subscript𝒬𝑖1subscript𝒬𝑖12subscript𝒬𝑖\displaystyle+\frac{1}{\lambda^{2}}(\vec{\mathcal{Q}}_{i-1}+\vec{\mathcal{Q}}_{i+1}-2\vec{\mathcal{Q}}_{i}).

where λ>0𝜆0\lambda>0 is the coupling between lattice charges and hair, and σ𝜎\sigma determines the velocity of propagation of hair in the continuum limit. The first equation above gives the equations for evolution of ti(u)subscript𝑡𝑖𝑢t_{i}(u) and the second equation determines the evolution of the hair. These equations have unique solutions provided we specify the initial lattice charges, and the hair charges and their time-derivatives (we assume initial synchronicity, i.e. ti(u=u0)=u0subscript𝑡𝑖𝑢subscript𝑢0subscript𝑢0t_{i}(u=u_{0})=u_{0} at initial time u0subscript𝑢0u_{0} but this is not necessary) and can be solved numerically following PhysRevD.101.066001 . Generalizations of the above equations with higher derivative corrections and desired phenomenological features are possible but we do not discuss them here. These equations imply a specific form of null matter in the AdS2𝐴𝑑subscript𝑆2AdS_{2} throats, but alternatively we can simply think of these as the equations determining the quantum dots and hair. Crucially, in the full interacting system we have only one global SL(2,R)𝑆𝐿2𝑅SL(2,R) symmetry, namely that of the original unfragmented geometry.

Note that we should treat the lattice charges 𝒬isubscript𝒬𝑖\vec{\mathcal{Q}}_{i} semi-classically invoking a large-N𝑁N limit in each throat, but the hair 𝖰isubscript𝖰𝑖\vec{\mathsf{Q}}_{i} should be understood as an open quantum system interacting with the lattice. However, for present purposes it will be sufficient to consider coherent states of the hair and treat it classically. The equations (5.3.2) imply that the full system has a conserved energy of the form:

=𝒬+𝖰subscript𝒬subscript𝖰\mathcal{E}=\mathcal{E}_{\mathcal{Q}}+\mathcal{E}_{\mathsf{Q}} (106)

which is simply a sum of the energy in the lattice charges and that in the hair with

𝒬subscript𝒬\displaystyle\mathcal{E}_{\mathcal{Q}} =\displaystyle= iMi=i𝒬i𝒬i,subscript𝑖subscript𝑀𝑖subscript𝑖subscript𝒬𝑖subscript𝒬𝑖\displaystyle\sum_{i}M_{i}=\sum_{i}\mathcal{Q}_{i}\cdot\mathcal{Q}_{i},
𝖰subscript𝖰\displaystyle\mathcal{E}_{\mathsf{Q}} =\displaystyle= λ32i𝖰i𝖰isuperscript𝜆32subscript𝑖superscriptsubscript𝖰𝑖superscriptsubscript𝖰𝑖\displaystyle\frac{\lambda^{3}}{2}\sum_{i}\vec{\mathsf{Q}}_{i}^{\prime}\cdot\vec{\mathsf{Q}}_{i}^{\prime} (107)
+λ32σ2i(𝖰i𝖰i1)(𝖰i𝖰i1).superscript𝜆32superscript𝜎2subscript𝑖subscript𝖰𝑖subscript𝖰𝑖1subscript𝖰𝑖subscript𝖰𝑖1\displaystyle+\frac{\lambda^{3}}{2\sigma^{2}}\sum_{i}(\vec{\mathsf{Q}}_{i}-\vec{\mathsf{Q}}_{i-1})\cdot(\vec{\mathsf{Q}}_{i}-\vec{\mathsf{Q}}_{i-1}).

Thus 𝒬subscript𝒬\mathcal{E}_{\mathcal{Q}} is simply the sum of the ADM masses of the throats and 𝖰subscript𝖰\mathcal{E}_{\mathsf{Q}} is the (discretised) kinetic energy of the hair. Note that λ>0𝜆0\lambda>0 is necessary for the positivity of the average energy in the microcanonical ensemble as we will see below.

The microstates: The microstates of the black hole can be identified with stationary solutions of (5.3.2). One can readily prove that in such microstates, we should have

𝒬i=Qξ+𝒬i,with𝒬iξ=0.formulae-sequencesubscript𝒬𝑖𝑄𝜉superscriptsubscript𝒬𝑖perpendicular-towithsuperscriptsubscript𝒬𝑖perpendicular-to𝜉0\vec{\mathcal{Q}}_{i}=Q\vec{\xi}+\vec{\mathcal{Q}}_{i}^{\perp},\quad{\rm with}\quad\vec{\mathcal{Q}}_{i}^{\perp}\cdot\xi=0. (108)

The SL(2,R)𝑆𝐿2𝑅SL(2,R) charge vector ξ𝜉\vec{\xi} thus spontaneously breaks the global SL(2,R)𝑆𝐿2𝑅SL(2,R) symmetry. We can set this global frame ξ𝜉\vec{\xi} in the 00-direction without loss of generality so that 𝒬isuperscriptsubscript𝒬𝑖perpendicular-to\vec{\mathcal{Q}}_{i}^{\perp} have only ++ and - components. We also normalize ξ𝜉\vec{\xi} such that ξξ=1𝜉𝜉1\vec{\xi}\cdot\vec{\xi}=1. Furthermore, for stationarity we need the hair charges to have the following configuration:

𝖰i(u)=𝖰iloc+𝖰imon(u)+qirad(u)ξsubscript𝖰𝑖𝑢superscriptsubscript𝖰𝑖locsuperscriptsubscript𝖰𝑖mon𝑢superscriptsubscript𝑞𝑖rad𝑢𝜉\vec{\mathsf{Q}}_{i}(u)=\vec{\mathsf{Q}}_{i}^{\rm loc}+\vec{\mathsf{Q}}_{i}^{\rm mon}(u)+q_{i}^{\rm rad}(u)\vec{\xi} (109)

where

𝖰iloc=σ2λ2𝒬i+𝒦,superscriptsubscript𝖰𝑖locsuperscript𝜎2superscript𝜆2subscript𝒬𝑖𝒦\vec{\mathsf{Q}}_{i}^{\rm loc}=-\frac{\sigma^{2}}{\lambda^{2}}\vec{\mathcal{Q}}_{i}+\vec{\mathcal{K}}, (110)

are locked to the lattice charges,

𝖰imon=αξusuperscriptsubscript𝖰𝑖mon𝛼𝜉𝑢\vec{\mathsf{Q}}_{i}^{\rm mon}=\alpha\vec{\xi}u (111)

is a monopole component with a homogeneous 𝖰i=αξsuperscriptsubscript𝖰𝑖𝛼𝜉\vec{\mathsf{Q}}_{i}^{\prime}=\alpha\xi in the direction of the global frame, and qirad(u)superscriptsubscript𝑞𝑖rad𝑢q_{i}^{\rm rad}(u) are oscillating hair components decoupled from the lattice and satisfying the normal mode equations

qirad′′=1σ2(qi1rad+qi+1rad2qirad).superscriptsuperscriptsubscript𝑞𝑖rad′′1superscript𝜎2superscriptsubscript𝑞𝑖1radsuperscriptsubscript𝑞𝑖1rad2superscriptsubscript𝑞𝑖rad{q_{i}^{\rm rad}}^{\prime\prime}=\frac{1}{\sigma^{2}}(q_{i-1}^{\rm rad}+q_{i+1}^{\rm rad}-2q_{i}^{\rm rad}). (112)

To avoid redundancy, we can set

iqirad=iqirad=0.subscript𝑖superscriptsubscript𝑞𝑖radsubscript𝑖superscriptsuperscriptsubscript𝑞𝑖rad0\sum_{i}q_{i}^{\rm rad}=\sum_{i}{q_{i}^{\rm rad}}^{\prime}=0. (113)

Crucially the Fourier transform of qirad(u)superscriptsubscript𝑞𝑖rad𝑢q_{i}^{\rm rad}(u) will have support only on the discrete normal modes with non-vanishing frequencies.

It follows from (5.3.2) that i𝖰isubscript𝑖superscriptsubscript𝖰𝑖\sum_{i}\vec{\mathsf{Q}_{i}^{\prime}} is conserved, and therefore α𝛼\alpha, the monopole charge, should not change when the microstate is perturbed.

Finally, we can define the microcanonical ensemble as the collection of microstate solutions subject to the constraints, that (i) the total Misubscript𝑀𝑖M_{i} should add up to the ADM mass of the black hole, i.e. 𝒬=Msubscript𝒬𝑀\mathcal{E}_{\mathcal{Q}}=M , and (ii) ti(u)subscript𝑡𝑖𝑢t_{i}(u) and hence τi(u)subscript𝜏𝑖𝑢\tau_{i}(u) should be real, continuous and have continuous first and second derivatives at all lattice sites (necessary to define the mass Misubscript𝑀𝑖M_{i} that is proportional to the Schwarzian derivative). This implies two set of possibilities. Firstly let’s set ξ0=1superscript𝜉01\xi^{0}=1 and ξ±=0superscript𝜉plus-or-minus0\xi^{\pm}=0 without loss of generality as mentioned before. Then the first set of possibilities which satisfy both constraints are those which satisfy 0MiQ20subscript𝑀𝑖superscript𝑄20\leq M_{i}\leq Q^{2}, Q>0𝑄0Q>0, iMi=Msubscript𝑖subscript𝑀𝑖𝑀\sum_{i}M_{i}=M and

𝒬i0=Q,𝒬i+=ρiQ2M,formulae-sequencesuperscriptsubscript𝒬𝑖0𝑄superscriptsubscript𝒬𝑖subscript𝜌𝑖superscript𝑄2𝑀\displaystyle\mathcal{Q}_{i}^{0}=-Q,\quad\mathcal{Q}_{i}^{+}=-\rho_{i}\sqrt{Q^{2}-M},
𝒬i=1ρiQ2Msuperscriptsubscript𝒬𝑖1subscript𝜌𝑖superscript𝑄2𝑀\displaystyle\mathcal{Q}_{i}^{-}=-\frac{1}{\rho_{i}}\sqrt{Q^{2}-M} (114)

with

QMiQ+MiρiQ+MiQMi.𝑄subscript𝑀𝑖𝑄subscript𝑀𝑖subscript𝜌𝑖𝑄subscript𝑀𝑖𝑄subscript𝑀𝑖\displaystyle\sqrt{\frac{Q-\sqrt{M_{i}}}{Q+\sqrt{M_{i}}}}\leq\rho_{i}\leq\sqrt{\frac{Q+\sqrt{M_{i}}}{Q-\sqrt{M_{i}}}}. (115)

Remarkably, these imply that ti0superscriptsubscript𝑡𝑖0t_{i}^{\prime}\geq 0 (and hence τi0superscriptsubscript𝜏𝑖0\tau_{i}^{\prime}\geq 0), i.e. the arrows of time of all the lattice sites should be aligned towards the future. Thus a global arrow of time is a consequence of the equations of motion! The second set of possibilities lead to similar inequalities which align the global arrow of time towards the past. We discard this set and define the microcanonical ensemble with the mentioned set of (in)equalities.

In the microstate solutions, the total energy of the hair 𝖰subscript𝖰\mathcal{E}_{\mathsf{Q}} further splits into three parts, i.e.

𝖰=𝖰pot+𝖰mon+𝖰rad.subscript𝖰superscriptsubscript𝖰potsuperscriptsubscript𝖰monsuperscriptsubscript𝖰rad\mathcal{E}_{\mathsf{Q}}=\mathcal{E}_{\mathsf{Q}}^{\rm pot}+\mathcal{E}_{\mathsf{Q}}^{\rm mon}+\mathcal{E}_{\mathsf{Q}}^{\rm rad}. (116)

with (setting ξ0=1superscript𝜉01\xi^{0}=1 and ξ±=0superscript𝜉plus-or-minus0\xi^{\pm}=0 without loss of generality as mentioned before )

𝖰potsuperscriptsubscript𝖰pot\displaystyle\mathcal{E}_{\mathsf{Q}}^{\rm pot} =\displaystyle= σ22λi(𝒬i+𝒬i1+)(𝒬i𝒬i),superscript𝜎22𝜆subscript𝑖superscriptsubscript𝒬𝑖superscriptsubscript𝒬𝑖1superscriptsubscript𝒬𝑖superscriptsubscript𝒬𝑖\displaystyle-\frac{\sigma^{2}}{2\lambda}\sum_{i}(\mathcal{Q}_{i}^{+}-\mathcal{Q}_{i-1}^{+})(\mathcal{Q}_{i}^{-}-\mathcal{Q}_{i}^{-}),
𝖰monsuperscriptsubscript𝖰mon\displaystyle\mathcal{E}_{\mathsf{Q}}^{\rm mon} =\displaystyle= 12λ3α2,12superscript𝜆3superscript𝛼2\displaystyle\frac{1}{2}\lambda^{3}\alpha^{2},
𝖰radsuperscriptsubscript𝖰rad\displaystyle\mathcal{E}_{\mathsf{Q}}^{\rm rad} =\displaystyle= λ32iqirad2+λ32σ2i(qiqi1)2.superscript𝜆32subscript𝑖superscriptsuperscriptsuperscriptsubscript𝑞𝑖rad2superscript𝜆32superscript𝜎2subscript𝑖superscriptsubscript𝑞𝑖subscript𝑞𝑖12\displaystyle\frac{\lambda^{3}}{2}\sum_{i}{{q_{i}^{\rm rad}}^{\prime}}^{2}+\frac{\lambda^{3}}{2\sigma^{2}}\sum_{i}(q_{i}-q_{i-1})^{2}. (117)

Clearly, if λ>0𝜆0\lambda>0, then both 𝖰monsuperscriptsubscript𝖰mon\mathcal{E}_{\mathsf{Q}}^{\rm mon} and 𝖰redsuperscriptsubscript𝖰red\mathcal{E}_{\mathsf{Q}}^{\rm red} are positive. Although 𝖰potsuperscriptsubscript𝖰pot\mathcal{E}_{\mathsf{Q}}^{\rm pot} need not be positive, its ensemble average is zero. Thus the average energy is positive if λ>0𝜆0\lambda>0.

Both classically and quantum-mechanically, the hair cannot be separated into interior and exterior components. However, in microstate solutions such a split operationally emerges since 𝖰ilocsuperscriptsubscript𝖰𝑖loc\vec{\mathsf{Q}}_{i}^{\rm loc} is locked with the lattice SL(2,R)𝑆𝐿2𝑅SL(2,R) charges which describe the configuration of the black hole interior, while both 𝖰imonsuperscriptsubscript𝖰𝑖mon\vec{\mathsf{Q}}_{i}^{\rm mon} and 𝖰iradsuperscriptsubscript𝖰𝑖rad\vec{\mathsf{Q}}_{i}^{\rm rad} do not affect the interior and are thus decoupled from it. The potential energy term 𝖰potsuperscriptsubscript𝖰pot\mathcal{E}_{\mathsf{Q}}^{\rm pot} is determined solely by 𝖰ilocsuperscriptsubscript𝖰𝑖loc\vec{\mathsf{Q}}_{i}^{\rm loc}, while 𝖰monsuperscriptsubscript𝖰mon\mathcal{E}_{\mathsf{Q}}^{\rm mon} and 𝖰radsuperscriptsubscript𝖰rad\mathcal{E}_{\mathsf{Q}}^{\rm rad} are determined by 𝖰imonsuperscriptsubscript𝖰𝑖mon\vec{\mathsf{Q}}_{i}^{\rm mon} and 𝖰iradsuperscriptsubscript𝖰𝑖rad\vec{\mathsf{Q}}_{i}^{\rm rad} respectively. Therefore, the total energy also splits into an interior component which is the sum of 𝒬=Msubscript𝒬𝑀\mathcal{E}_{\mathcal{Q}}=M and 𝖰potsuperscriptsubscript𝖰pot\mathcal{E}_{\mathsf{Q}}^{\rm pot}, and the exterior component which is the sum of 𝖰imonsuperscriptsubscript𝖰𝑖mon\vec{\mathsf{Q}}_{i}^{\rm mon} and 𝖰iradsuperscriptsubscript𝖰𝑖rad\vec{\mathsf{Q}}_{i}^{\rm rad}. Operationally, therefore, the Hilbert space of the hair has the structure

𝖰=α𝖰αint𝖰αextsubscript𝖰subscriptdirect-sum𝛼tensor-productsuperscriptsubscriptsubscript𝖰𝛼intsuperscriptsubscriptsubscript𝖰𝛼ext\mathcal{H}_{\mathsf{Q}}=\bigoplus_{\alpha}\mathcal{H}_{\mathsf{Q}_{\alpha}}^{\rm int}\otimes\mathcal{H}_{\mathsf{Q}_{\alpha}}^{\rm ext}

with α𝛼\alpha denoting microstates assuming that the evolution is adiabatic (we can employ an adiabatically evolving basis) with the off-diagonal terms being suppressed due to decoherence. Note that this would imply that the global frame ξ𝜉\vec{\xi} which is crucial to make the distinction between the interior and exterior would be evolving adiabatically too. We would come back to this in the context of the Hawking evaporation with asymptotic boundary conditions at the throats that allow the Hawking quanta to escape.

Phenomenological viability: The immediate question is that whether the ensemble of microstates in the model behave like a semi-classical black hole with relaxing and energy-absorbing properties. To see this, we once again return to the semi-classical limit in which Hawking radiation is absent, and perturb an arbitrary microstate solution in the ensemble by a sequence of shocks (injections of energies) eiAsubscript𝑒𝑖𝐴e_{iA} with i𝑖i denoting the i𝑖i-th throat and A𝐴A referring to the instant uAsubscript𝑢𝐴u_{A}. The equations in (5.3.2) are then modified to

Misuperscriptsubscript𝑀𝑖\displaystyle M_{i}^{\prime} =\displaystyle= λ(𝒬i1+𝒬i+12𝒬i)𝖰i,𝜆subscript𝒬𝑖1subscript𝒬𝑖12subscript𝒬𝑖superscriptsubscript𝖰𝑖\displaystyle-\lambda(\vec{\mathcal{Q}}_{i-1}+\vec{\mathcal{Q}}_{i+1}-2\vec{\mathcal{Q}}_{i})\cdot\vec{\mathsf{Q}}_{i}^{\prime},
+AeiAδ(uuA),subscript𝐴subscript𝑒𝑖𝐴𝛿𝑢subscript𝑢𝐴\displaystyle+\sum_{A}e_{iA}\delta(u-u_{A}),
𝖰i′′superscriptsubscript𝖰𝑖′′\displaystyle\vec{\mathsf{Q}}_{i}^{\prime\prime} =\displaystyle= 1σ2(𝖰i1+𝖰i+12𝖰i)1superscript𝜎2subscript𝖰𝑖1subscript𝖰𝑖12subscript𝖰𝑖\displaystyle\frac{1}{\sigma^{2}}(\vec{\mathsf{Q}}_{i-1}+\vec{\mathsf{Q}}_{i+1}-2\vec{\mathsf{Q}}_{i}) (118)
+1λ2(𝒬i1+𝒬i+12𝒬i).1superscript𝜆2subscript𝒬𝑖1subscript𝒬𝑖12subscript𝒬𝑖\displaystyle+\frac{1}{\lambda^{2}}(\vec{\mathcal{Q}}_{i-1}+\vec{\mathcal{Q}}_{i+1}-2\vec{\mathcal{Q}}_{i}).

Note that the hair is not directly coupling to the shocks (this would imply the results such as mirroring to be more non-trivial). The shocks are localized on the lattice and cannot directly affect the delocalized gravitational charges comprising the hair. The shocks add a total energy of

Δ=i,Aei,AΔsubscript𝑖𝐴subscript𝑒𝑖𝐴\Delta\mathcal{E}=\sum_{i,A}e_{i,A}

to the system. Starting from a random microstate with the (conserved) monopole charge α>0𝛼0\alpha>0 and simulating the system via methods of PhysRevD.101.066001 , one finds that

  1. 1.

    Any microstate with or without decoupled hair oscillations rapidly settles down to another microstate with decoupled hair oscillations after the sequence of shocks.

  2. 2.

    The final microstate is determined by the initial microstate in a rather complex manner. Even for a single shock, one needs to take into account the initial states of all lattice sites and that of the hair indicating pseudorandom dynamics.

  3. 3.

    Almost all the energy in the shock is absorbed by the change in the total black hole mass, i.e. 𝒬=Msubscript𝒬𝑀\mathcal{E}_{\mathcal{Q}}=M. This statement becomes better (for a typical initial microstate) if we increase the number of sites keeping the total initial black hole mass and total initial energy fixed. However, even for five lattice sites, less than 111 percent of the energy in shocks is transferred to the hair.

See Fig. 28 for an illustration for the case of a five site lattice suffering a single shock. The above features imply that a typical microstate with a positive monopole charge behaves like a semi-classical black hole qualitatively as far its relaxation dynamics and energy absorption properties are concerned, and furthermore shows features of pseudorandom dynamics.

Refer to caption
(a) Evolution of lattice charges and masses after a single shock
Refer to caption
(b) Evolution of the energies after a single shock
Figure 28: The evolution of a randomly chosen microstate after a single shock in a five site model is shown above. The first site is shocked with energy injection e=0.4𝑒0.4e=0.4 at u=0𝑢0u=0. We set λ=1𝜆1\lambda=1, α=1𝛼1\alpha=1 and σ=0.01𝜎0.01\sigma=0.01. The system relaxes to another microstate with a different distribution of masses and lattice charges. Almost the entire energy of the shock is absorbed by the total mass of the black hole 𝒬subscript𝒬\mathcal{E}_{\mathcal{Q}}. The energy in the hair 𝖰subscript𝖰\mathcal{E}_{\mathsf{Q}} remains approximately constant. These figures are from PhysRevD.102.086008 .

As a consequence of these phenomenological features, we can argue that the effective split into interior and exterior emerges dynamically after the relaxation time (which goes to a finite value in the continuum limit as observed numerically).272727 𝖰intsuperscriptsubscript𝖰int\mathcal{H}_{\mathsf{Q}}^{\rm int} is supported at the inhomogeneous static fixed point configurations (or analogous adiabatic versions of these in the presence of Hawking radiation) while 𝖰extsuperscriptsubscript𝖰ext\mathcal{H}_{\mathsf{Q}}^{\rm ext} is supported at the normal mode frequencies only (or analogous adiabatic versions of these in the presence of Hawking radiation). In very general situations, we expect such a split to be operationally valid when the dynamics is coarse-grained over the relaxation time-scale, and can be demonstrated by employing a suitable coherent state basis.

Classical information mirroring in hair: The remarkable aspect of these microstate models is that it has the feature of information mirroring built into them intrinsically even though the Hawking radiation is arrested via boundary conditions at the throats. This can be demonstrated readily by studying the dynamics of the system in response to a sequence of shocks. As described above, a typical microstate relaxes to another microstate demonstrating pseudorandom dynamics. A part of the hair 𝖰iradsuperscriptsubscript𝖰𝑖rad\vec{\mathsf{Q}}_{i}^{\rm rad} decouples from the interior while another component gets locked to the interior lattice charges. The information of the shocks is mirrored in the 𝖰iradsuperscriptsubscript𝖰𝑖rad\vec{\mathsf{Q}}_{i}^{\rm rad} or equivalently in qiradsuperscriptsubscript𝑞𝑖radq_{i}^{\rm rad} defined in (109). The crucial point is that there are features in qiradsuperscriptsubscript𝑞𝑖radq_{i}^{\rm rad} which allows us to decode the information of the shocks without the knowledge of the interior of the initial or final microstate. Many aspects of the information encoded in the infalling shocks can thus be decoded by assuming that the interior is in the microcanocial ensemble. The only necessary information for the decoding is the global frame ξ𝜉\xi in (109) which one can simply obtain from the (decoupled) monopole component, the analogue of the early radiation in the Hayden-Preskill thought experiment.

The Fourier transform of qirad(u)superscriptsubscript𝑞𝑖rad𝑢q_{i}^{\rm rad}(u) will be supported only at the non-vanishing normal mode frequencies of (112). It is sufficient to look into the phases and amplitudes of the positive normal mode frequencies since qirad(u)superscriptsubscript𝑞𝑖rad𝑢q_{i}^{\rm rad}(u) is real. It turns out that only the differences of the phases of the two positive normal mode frequencies at the various sites are necessary to decode the information of the shocks. If we encode information into the time sequence of the location of sites which are shocked and the differences of energy injections into these shocks are not large, then the information of the sequence can be decoded from the ordering of the phase differences of the positive frequency normal modes of qirad(u)superscriptsubscript𝑞𝑖rad𝑢q_{i}^{\rm rad}(u) at various lattice sites. All other features of qirad(u)superscriptsubscript𝑞𝑖rad𝑢q_{i}^{\rm rad}(u) will be contaminated with the details of the initial microstate but this ordering of phase differences will be determined only by the specific time sequence of the shocks.

For a non-trivial example, consider the case of five sites and a sequence of two shocks into two sites. The decoding protocol for which sites have been shocked and in which sequence is illustrated in Fig. 29. This protocol is independent of the initial microstate.

Refer to caption
(a) A sequence of two shocks into two non-neighboring sites. The dark blue site is shocked first and the light blue one later. The one shocked earlier has the highest and the one shocked later has the lowest phase differences between the positive normal modes.
Refer to caption
(b) A sequence of two shocks into two neighboring sites. The dark blue site is shocked first and the light blue one later. The one shocked earlier has the lowest and the one shocked later has the highest phase differences between the positive normal modes.
Refer to caption
(c) A sequence of two shocks into the same site. The shocked site has the highest phase differences between the positive normal modes.
Figure 29: The decoding protocol for information mirroring when two sites are shocked in the five site microstate model via measurements of the phase differences between the two positive normal modes in the hair oscillations which decouple from the interior. Note that this protocol for decoding which sites have been shocked and in which sequence is the same for any initial microstate. Although the absolute values of the phase differences are somewhat sensitive to the initial microstate, the ordering is not.

The information mirroring should carry over to the quantum regime if we quantize the hair. It would be interesting to find out if the information of the initial microstate prior to the shock is encoded in observables which are complementary, i.e. have large non-vanishing commutators with the phase differences in the normal modes. This should have consequences for understanding the black hole complementarity as we comment below.

Encoding into the Hawking radiation: Hawking radiation can be readily incorporated into the microstate model described above by simply coupling each throat to a bath and implementing transparent boundary conditions following AEMM as for instance. These boundary conditions choose a specific global null frame ξ𝜉\vec{\xi} (as for instance in AEMM there is a null shock when the boundary conditions are altered). The first question to be asked is whether the information of the interior leaks out to the Hawking radiation completely. It is a non-trivial question because although the two dimensional black holes evaporate away at all the lattice sites, the SL(2,R)𝑆𝐿2𝑅SL(2,R) directions of the lattice charges 𝒬isubscript𝒬𝑖\vec{\mathcal{Q}}_{i} may not homogenize and thus there would be a remnant information in the interior even after complete evaporation. One may need this remnant to decode the complete information. Firstly, we see that the SL(2,R)𝑆𝐿2𝑅SL(2,R) directions of the lattice charges 𝒬isubscript𝒬𝑖\vec{\mathcal{Q}}_{i} align themselves with the null direction ξ𝜉\vec{\xi} set by the boundary conditions asymptotically at late time. Nevertheless, asymptotically 𝒬i𝒬iξsubscript𝒬𝑖subscript𝒬𝑖𝜉\vec{\mathcal{Q}}_{i}\rightarrow\mathcal{Q}_{i}\vec{\xi}, so that the overall magnitude 𝒬isubscript𝒬𝑖\mathcal{Q}_{i} is non-trivial although the Casimir (the mass of the black hole) vanishes at each site. However, the hair (being a smaller subsystem) will also dissipate its energy to the Hawking radiation that escapes away. The evolution equations (5.3.2) would imply that the hair charges should lock themselves with the lattice charges 𝒬i𝒬iξsubscript𝒬𝑖subscript𝒬𝑖𝜉\vec{\mathcal{Q}}_{i}\approx\mathcal{Q}_{i}\vec{\xi} following (110) so that the potential energy vanishes (since ξ𝜉\vec{\xi} is null) and thus the hair system reaches a fixed point. Furthermore, as discussed below the hair should be entangled with the Hawking radiation so the full decoding of the interior should be possible.

The AMPS paradox, of course, is not about the final state after complete evaporation, but the system after Page time. Here, each individual throat would have its own Page time and therefore the Page time here refers to the average Page time of these throats (which will have small fluctuation about the average for a typical microstate of a large black hole). As argued above, the full system should have an effective split into interior and exterior because the hair can be split into 𝖰=α𝖰αint𝖰αextsubscript𝖰subscriptdirect-sum𝛼tensor-productsuperscriptsubscriptsubscript𝖰𝛼intsuperscriptsubscriptsubscript𝖰𝛼ext\mathcal{H}_{\mathsf{Q}}=\bigoplus_{\alpha}\mathcal{H}_{\mathsf{Q}_{\alpha}}^{\rm int}\otimes\mathcal{H}_{\mathsf{Q_{\alpha}}}^{\rm ext} when averaged over the (slowly evolving) relaxation time with α𝛼\alpha denoting different microstates. Also the outgoing Hawking radiation essentially forms a sequence of outgoing shocks. Therefore the information of coarse grained features of the outgoing Hawking radiation such as the overall energy outflow from the throats over a period commensurate with the relaxation time (which are insufficient to fully reconstruct the interior, especially the SL(2,R)𝑆𝐿2𝑅SL(2,R) directions of the lattice charges) will be encoded into the radiation component of the hair which essentially forms 𝖰extsuperscriptsubscript𝖰ext\mathcal{H}_{\mathsf{Q}}^{\rm ext} as happened in the case of mirroring of the information of the shocks. Crucially, these low energy/ coarse-grained features will be encoded in special observables such as the phase differences of the normal modes in 𝖰extsuperscriptsubscript𝖰ext\mathcal{H}_{\mathsf{Q}}^{\rm ext} from which decoding will be possible without the knowledge of the finer structure of the microstates. The detailed knowledge of the interior would be encoded in the complementary observables of 𝖰extsuperscriptsubscript𝖰ext\mathcal{H}_{\mathsf{Q}}^{\rm ext} and also 𝖰intsuperscriptsubscript𝖰int\mathcal{H}_{\mathsf{Q}}^{\rm int}282828In Verlinde:2012cy , one can find a general discussion on how effective field theory observables in a semiclassical but not necessarily smooth horizon geometry can decode the black hole interior. This discussion, cast in the formalism of quantum error correction, can be applied in the context of our microstate model. The crucial part of the construction of the recovery map from the local operators acting on the hair and the Hawking quanta at the horizon, is the knowledge of conditional transition matrices of the black hole interior. The resolution of the AMPS paradox in this context has been also discussed in Verlinde:2012cy ..

Finally to understand encoding in the Hawking radiation, we should decipher which observables acting on the Hawking radiation will be strongly correlated with features of 𝖰extsuperscriptsubscript𝖰ext\mathcal{H}_{\mathsf{Q}}^{\rm ext} that are responsible for mirroring the infalling qubits and encoding the coarse-grained information carried out by the outgoing Hawking radiation (these should have simple encoding in the sense that decoding can be possible by assuming that the interior is a microcanonical ensemble), and similarly which features of the Hawking radiation correlate strongly with 𝖰intsuperscriptsubscript𝖰int\mathcal{H}_{\mathsf{Q}}^{\rm int} and the fine grained features of the black hole interior (eg. the lattice SL(2,R)𝑆𝐿2𝑅SL(2,R) charges). The encoding of the latter should be complex due to the underlying pseudorandom dynamics of the microstates (more on the generation of complexity via the Python lunch mechanism below). Since the split between the interior and the exterior is only an operational concept that emerges after averaging over an evolving relaxation timescale, there is no real contradiction that the Hawking radiation is in a way entangled maximally with both 𝖰extsuperscriptsubscript𝖰ext\mathcal{H}_{\mathsf{Q}}^{\rm ext} and 𝖰intsuperscriptsubscript𝖰int\mathcal{H}_{\mathsf{Q}}^{\rm int}. Furthermore, the coarse-grained and finer information could be encoded with the best fidelity in observables that have large mutual commutators. This could be fundamental to understanding why the semiclassical geometry of the black hole emerges. Additionally, one possible way to realize the expected near saturation (but not violation) of the strong subadditivity property which is key to the AMPS paradox is that the tripartite interior, hair and outgoing Hawking radiation system has a quantum Markov chain like structure (see Section 4.2.2 for the definition of quantum Markov chain states).

The key to the detailed understanding would be to follow the evolution of the full tripartite system (after quantizing the hair) and analyze it using various tools of quantum information theory. This should reveal how black hole complementarity emerges in an operational sense free of paradoxes along with the self-averaging properties represented by the semi-classical black hole geometry and the Euclidean replica wormhole saddles.

Pseudorandom dynamics crucial for Python lunches: The appearance of Python lunch geometries gives a natural quantification of encoding of the interior modes into outgoing Hawking radiation as shown in Sec. 5.2. As discussed there, instead of the vacuum of the bulk matter in each throat, we should consider first a small subspace of its Hilbert space which we can identify with the code subspace, and then study the microstate dynamics with the bulk matter in the throat in the maximally mixed state in this (small) code subspace (the importance of the maximally mixed state in the code subspace has been emphasized also). Via the arguments presented in Sec. 5.2 we expect nucleation of a Python lunch or equivalently a (non-classical) locally maximal extremal surface in the throat. However, the inherent pseudorandom dynamics would be needed to ensure a necessary amplification so that the Python’s lunch manifests macroscopically across a significant fraction of throats (this would be needed in the higher dimensional setup) before the operational split of the hair Hilbert space into interior and exterior emerges, i.e. before black hole complementarity becomes operational. One can also similarly consider a code subspace of the interior modes of the hair (modes with frequencies smaller than the first non-trivial normal mode) and study the Python lunch phenomenon. It will be also interesting to consider code subspaces of interior modes of bulk matter involving entanglement across several throats since it will give insights into understanding of the complexity of encoding of the entanglement of the interior. In all cases, the inherent chaos in microstate dynamics should connect the Python’s lunch geometric mechanism of generating exponential complexity Brown:2019rox in encoding with the pseudorandomness mechanism discussed in Kim:2020cds .

Outlook: These microstate models hold many promises for partial answers to how the encoding in Hawking radiation happens with desired features along with the validation of black hole complementarity, but further improvements would be necessary. Although we find information mirroring occurring at relaxation time in the microstates, it is not clear if they have shortest possible scrambling time for typical microstates. Furthermore, the bath region which collects the Hawking radiation is itself fragmented into two dimensional spaces. It would be necessary to glue these baths in a way we can form a connected flat asymptotic spacetime region. One way to construct more realistic models would be to figure out how simplified scenarios with characteristics of those discussed above can emerge from fuzzballs in string theory. The notion of fuzzball complementarity Mathur:2010kx ; Avery:2012tf which advocate emergent holographic descriptions for local measurements by an infalling observer could be useful.

6 Discussion and Outlook

One of the topics we have not been been able to discuss in this review is the fundamental relationship between tensor networks and gravity. Although tensor networks do model many aspects of holography, it is unclear if they reproduce all information theoretic aspects beyond the Ryu-Takanayagi type extremal surface which computes entanglement entropy. The latter was established first Swingle:2009bg in the context of multiscale-entanglement-renormalization (MERA) tensor networks.292929The correspondence between tensor networks and holography has also been explored via continuum versions of MERA Miyaji:2015yva . See also Caputa:2017urj for a path integral formulation in which the continuum version of the tensor network is reformulated as an optimization problem. In Akers:2018fow it has been argued that existing quantum error correcting tensor networks in the toy models discussed in section 4.2.3 are area eigenstates of the bulk gravity theory rather than representatives of smooth bulk geometries. A better understanding of these issues could also come from constructing tensor networks which mimic the highly efficient RG flow described in section 4.4 as it is designed to reproduce dynamical gravity. These issues are important to understand how we can simulate aspects of quantum gravity and also strongly interacting quantum field theories in an efficient qubit regularized way and in real time.

Another interesting direction of research would be related to further elucidating the models of evaporating black holes in terms of quantum thermodynamics. These models described in this review involve coupling of a bath to a holographic system which has quantum matter in the bulk, and therefore usual ubiquitous outcomes like monotonic growth of entropy in classical gravity coupled to matter which satisfies the null energy condition, can be avoided. It could be possible to construct quantum engines in similar setups. Quantum thermodynamics is a rapidly emerging field which uses quantum information to generalize the laws of thermodynamics Goold_2016 ; PhysRevE.93.022126 ; Yunger_Halpern_2016 ; Guryanova_2016 ; RevModPhys.91.025001 . As for instance, bounds have been established for the one-shot work cost of creating a state and also for extractable work from the state in terms of the hypothesis testing relative entropy with respect to the Gibbs ensemble PhysRevE.93.022126 . Recently, even within the classical approximation in the bulk (with classical bulk matter) and assuming an infinite memoryless bath, it has been shown in Kibe:2021qjy that for an instantaneous transition between thermal rotating states in holography, the quantum null energy condition Bousso_2016 of the boundary theory bounds the growth of thermodynamic entropy (temperature) for a fixed increase in temperature (entropy) from both above and below. Similarly, the rates of growth of entanglement can also be bounded from both above and below. Furthermore, one can recover the Landauer erasure principle and also understand how to construct erasure tolerant quantum memory Banerjee:2022dgv . It would be interesting to pursue how one can construct various protocols by exploiting suitable quantum bulk matter in a semi-classical black hole geometry and interacting with a dynamical reservoir at the boundary. Also how the efficiency and work cost of such processes can be bounded via tools of quantum information theory especially the quantum null energy condition and its possible generalizations.

It is quite likely that the fundamental understanding bulk emergence and quantum black holes in holography will lead us to more fundamental and novel connections between quantum information and many-body dynamics.

Furthermore, we have restricted ourselves here to asymptotically anti-de Sitter space. For interesting discussions especially on how islands (entanglement wedges) may or may not generalize consistently for other asymptotic boundary conditions see Geng:2020fxl ; Geng:2021hlu ; Shaghoulian:2021cef (an explicit discussion on de-Sitter space is in Shaghoulian:2021cef ). In the future, it would be of deep interest to understand bulk emergence in the context of both asymptotically flat and cosmological spacetimes. In the latter context, the approach of Bzowski:2012ih , as for instance, may lead to promising results by extrapolating our existing understanding of anti-de Sitter spaces after incorporating extremal surfaces and consistent entanglement wedges.

Acknowledgements.
The research of TK is supported by the Prime Minister’s Research Fellowship (PMRF). PM and AM ackowledge support from the Institute of Eminence scheme of IIT Madras funded by the Ministry of Education of India. AM acknowledges the support of the Ramanujan Fellowship of the Science and Engineering Board of the Department of Science and Technology of India, the new faculty seed grant of IIT Madras and the CEFIPRA/IFCPAR grant no 6304-3.

References

  • (1) J.M. Maldacena, The Large N limit of superconformal field theories and supergravity, Adv. Theor. Math. Phys. 2, 231 (1998). DOI 10.1023/A:1026654312961
  • (2) S.S. Gubser, I.R. Klebanov, A.M. Polyakov, Gauge theory correlators from noncritical string theory, Phys. Lett. B 428, 105 (1998). DOI 10.1016/S0370-2693(98)00377-3
  • (3) E. Witten, Anti-de Sitter space and holography, Adv. Theor. Math. Phys. 2, 253 (1998). DOI 10.4310/ATMP.1998.v2.n2.a2
  • (4) L. Susskind, The World as a hologram, J. Math. Phys. 36, 6377 (1995). DOI 10.1063/1.531249
  • (5) J.D. Bekenstein, Black holes and entropy, Phys. Rev. D 7, 2333 (1973). DOI 10.1103/PhysRevD.7.2333
  • (6) J.M. Bardeen, B. Carter, S.W. Hawking, The Four laws of black hole mechanics, Commun. Math. Phys. 31, 161 (1973). DOI 10.1007/BF01645742
  • (7) J.D. Bekenstein, Generalized second law of thermodynamics in black hole physics, Phys. Rev. D 9, 3292 (1974). DOI 10.1103/PhysRevD.9.3292
  • (8) J.D. Bekenstein, Universal upper bound on the entropy-to-energy ratio for bounded systems, Phys. Rev. D 23, 287 (1981). DOI 10.1103/PhysRevD.23.287. URL https://link.aps.org/doi/10.1103/PhysRevD.23.287
  • (9) R. Bousso, Holography in general space-times, JHEP 06, 028 (1999). DOI 10.1088/1126-6708/1999/06/028
  • (10) R. Bousso, A Covariant entropy conjecture, JHEP 07, 004 (1999). DOI 10.1088/1126-6708/1999/07/004
  • (11) H. Casini, Relative entropy and the Bekenstein bound, Class. Quant. Grav. 25, 205021 (2008). DOI 10.1088/0264-9381/25/20/205021
  • (12) P. Banerjee. Some basics of AdS/CFT. https://www.imsc.res.in/~pinakib/AdS-CFT_ST4.pdf (2017). [Online]
  • (13) O. Aharony, S.S. Gubser, J.M. Maldacena, H. Ooguri, Y. Oz, Large N field theories, string theory and gravity, Phys. Rept. 323, 183 (2000). DOI 10.1016/S0370-1573(99)00083-6
  • (14) D. Mateos, String Theory and Quantum Chromodynamics, Class. Quant. Grav. 24, S713 (2007). DOI 10.1088/0264-9381/24/21/S01
  • (15) S.A. Hartnoll, A. Lucas, S. Sachdev, Holographic quantum matter, arXiv 1612.07324 (2016)
  • (16) Y. Kim, I.J. Shin, T. Tsukioka, Holographic QCD: Past, Present, and Future, Prog. Part. Nucl. Phys. 68, 55 (2013). DOI 10.1016/j.ppnp.2012.09.002
  • (17) A. Rebhan, The Witten-Sakai-Sugimoto model: A brief review and some recent results, EPJ Web Conf. 95, 02005 (2015). DOI 10.1051/epjconf/20159502005
  • (18) J. Casalderrey-Solana, H. Liu, D. Mateos, K. Rajagopal, U.A. Wiedemann, Gauge/String Duality, Hot QCD and Heavy Ion Collisions (Cambridge University Press, 2014). DOI 10.1017/CBO9781139136747
  • (19) N. Beisert, et al., Review of AdS/CFT Integrability: An Overview, Lett. Math. Phys. 99, 3 (2012). DOI 10.1007/s11005-011-0529-2
  • (20) K. Zarembo, Localization and AdS/CFT Correspondence, J. Phys. A 50(44), 443011 (2017). DOI 10.1088/1751-8121/aa585b
  • (21) L. Eberhardt, M.R. Gaberdiel, R. Gopakumar, Deriving the AdS3/CFT2 correspondence, JHEP 02, 136 (2020). DOI 10.1007/JHEP02(2020)136
  • (22) M.R. Gaberdiel, R. Gopakumar, String Dual to Free N=4 Supersymmetric Yang-Mills Theory, Phys. Rev. Lett. 127(13), 131601 (2021). DOI 10.1103/PhysRevLett.127.131601
  • (23) S. Ryu, T. Takayanagi, Holographic derivation of entanglement entropy from the anti–de sitter space/conformal field theory correspondence, Phys. Rev. Lett. 96, 181602 (2006). DOI 10.1103/PhysRevLett.96.181602. URL https://link.aps.org/doi/10.1103/PhysRevLett.96.181602
  • (24) V.E. Hubeny, M. Rangamani, T. Takayanagi, A covariant holographic entanglement entropy proposal, Journal of High Energy Physics 2007(07), 062 (2007). DOI 10.1088/1126-6708/2007/07/062. URL https://doi.org/10.1088/1126-6708/2007/07/062
  • (25) N. Engelhardt, A.C. Wall, Quantum Extremal Surfaces: Holographic Entanglement Entropy beyond the Classical Regime, JHEP 01, 073 (2015). DOI 10.1007/JHEP01(2015)073
  • (26) B. Czech, J.L. Karczmarek, F. Nogueira, M. Van Raamsdonk, The gravity dual of a density matrix, Classical and Quantum Gravity 29(15), 155009 (2012). DOI 10.1088/0264-9381/29/15/155009. URL http://dx.doi.org/10.1088/0264-9381/29/15/155009
  • (27) A.C. Wall, Maximin Surfaces, and the Strong Subadditivity of the Covariant Holographic Entanglement Entropy, Class. Quant. Grav. 31(22), 225007 (2014). DOI 10.1088/0264-9381/31/22/225007
  • (28) T. Faulkner, A. Lewkowycz, J. Maldacena, Quantum corrections to holographic entanglement entropy, Journal of High Energy Physics 2013(11), 74 (2013). DOI 10.1007/JHEP11(2013)074. URL https://doi.org/10.1007/JHEP11(2013)074
  • (29) M. Headrick, V.E. Hubeny, A. Lawrence, M. Rangamani, Causality & holographic entanglement entropy, Journal of High Energy Physics 2014(12) (2014). DOI 10.1007/jhep12(2014)162. URL http://dx.doi.org/10.1007/JHEP12(2014)162
  • (30) D.L. Jafferis, A. Lewkowycz, J. Maldacena, S.J. Suh, Relative entropy equals bulk relative entropy, Journal of High Energy Physics 2016(6) (2016). DOI 10.1007/jhep06(2016)004. URL http://dx.doi.org/10.1007/JHEP06(2016)004
  • (31) T. Faulkner, A. Lewkowycz, Bulk locality from modular flow, Journal of High Energy Physics 2017(7) (2017). DOI 10.1007/jhep07(2017)151. URL http://dx.doi.org/10.1007/JHEP07(2017)151
  • (32) A. Almheiri, X. Dong, D. Harlow, Bulk locality and quantum error correction in ads/cft, Journal of High Energy Physics 2015(4) (2015). DOI 10.1007/jhep04(2015)163. URL http://dx.doi.org/10.1007/JHEP04(2015)163
  • (33) D. Harlow, The Ryu–Takayanagi Formula from Quantum Error Correction, Commun. Math. Phys. 354(3), 865 (2017). DOI 10.1007/s00220-017-2904-z
  • (34) X. Dong, D. Harlow, A.C. Wall, Reconstruction of bulk operators within the entanglement wedge in gauge-gravity duality, Phys. Rev. Lett. 117, 021601 (2016). DOI 10.1103/PhysRevLett.117.021601. URL https://link.aps.org/doi/10.1103/PhysRevLett.117.021601
  • (35) J. Cotler, P. Hayden, G. Penington, G. Salton, B. Swingle, M. Walter, Entanglement wedge reconstruction via universal recovery channels, Physical Review X 9(3), 031011 (2019)
  • (36) R. Bousso, Z. Fisher, S. Leichenauer, A.C. Wall, Quantum focusing conjecture, Physical Review D 93(6) (2016). DOI 10.1103/physrevd.93.064044. URL http://dx.doi.org/10.1103/PhysRevD.93.064044
  • (37) J. Koeller, S. Leichenauer, Holographic proof of the quantum null energy condition, Physical Review D 94(2) (2016). DOI 10.1103/physrevd.94.024026. URL http://dx.doi.org/10.1103/PhysRevD.94.024026
  • (38) S. Balakrishnan, T. Faulkner, Z.U. Khandker, H. Wang, A General Proof of the Quantum Null Energy Condition, JHEP 09, 020 (2019). DOI 10.1007/JHEP09(2019)020
  • (39) R. Bousso, Z. Fisher, J. Koeller, S. Leichenauer, A.C. Wall, Proof of the quantum null energy condition, Physical Review D 93(2) (2016). DOI 10.1103/physrevd.93.024017. URL http://dx.doi.org/10.1103/PhysRevD.93.024017
  • (40) T.A. Malik, R. Lopez-Mobilia, Proof of the quantum null energy condition for free fermionic field theories, Phys. Rev. D 101(6), 066028 (2020). DOI 10.1103/PhysRevD.101.066028
  • (41) D.Z. Freedman, S.S. Gubser, K. Pilch, N.P. Warner, Renormalization group flows from holography supersymmetry and a c theorem, Adv. Theor. Math. Phys. 3, 363 (1999). DOI 10.4310/ATMP.1999.v3.n2.a7
  • (42) R.C. Myers, A. Sinha, Seeing a c-theorem with holography, Phys. Rev. D 82, 046006 (2010). DOI 10.1103/PhysRevD.82.046006
  • (43) H. Casini, M. Huerta, On the RG running of the entanglement entropy of a circle, Phys. Rev. D 85, 125016 (2012). DOI 10.1103/PhysRevD.85.125016
  • (44) S.W. Hawking, Particle Creation by Black Holes, Commun. Math. Phys. 43, 199 (1975). DOI 10.1007/BF02345020. [Erratum: Commun.Math.Phys. 46, 206 (1976)]
  • (45) S.W. Hawking, Breakdown of Predictability in Gravitational Collapse, Phys. Rev. D 14, 2460 (1976). DOI 10.1103/PhysRevD.14.2460
  • (46) K. Ghosh, C. Krishnan, Dirichlet baths and the not-so-fine-grained Page curve, JHEP 08, 119 (2021). DOI 10.1007/JHEP08(2021)119
  • (47) D.N. Page, Information in black hole radiation, Physical Review Letters 71(23), 3743 (1993). DOI 10.1103/physrevlett.71.3743. URL http://dx.doi.org/10.1103/PhysRevLett.71.3743
  • (48) D.N. Page, Time dependence of hawking radiation entropy, Journal of Cosmology and Astroparticle Physics 2013(09), 028 (2013). DOI 10.1088/1475-7516/2013/09/028. URL http://dx.doi.org/10.1088/1475-7516/2013/09/028
  • (49) A. Almheiri, N. Engelhardt, D. Marolf, H. Maxfield, The entropy of bulk quantum fields and the entanglement wedge of an evaporating black hole, JHEP 12, 063 (2019). DOI 10.1007/JHEP12(2019)063
  • (50) G. Penington, Entanglement Wedge Reconstruction and the Information Paradox, JHEP 09, 002 (2020). DOI 10.1007/JHEP09(2020)002
  • (51) A. Almheiri, R. Mahajan, J. Maldacena, Y. Zhao, The Page curve of Hawking radiation from semiclassical geometry, JHEP 03, 149 (2020). DOI 10.1007/JHEP03(2020)149
  • (52) G. Penington, S.H. Shenker, D. Stanford, Z. Yang, Replica wormholes and the black hole interior, arXiv 1911.11977 (2019)
  • (53) A. Almheiri, T. Hartman, J. Maldacena, E. Shaghoulian, A. Tajdini, Replica wormholes and the entropy of hawking radiation, Journal of High Energy Physics 2020(5), 13 (2020). DOI 10.1007/JHEP05(2020)013. URL https://doi.org/10.1007/JHEP05(2020)013
  • (54) M. Rangamani, T. Takayanagi, Holographic Entanglement Entropy, vol. 931 (Springer, 2017). DOI 10.1007/978-3-319-52573-0
  • (55) S.D. Mathur, The Information paradox: A Pedagogical introduction, Class. Quant. Grav. 26, 224001 (2009). DOI 10.1088/0264-9381/26/22/224001
  • (56) D. Harlow, Jerusalem Lectures on Black Holes and Quantum Information, Rev. Mod. Phys. 88, 015002 (2016). DOI 10.1103/RevModPhys.88.015002
  • (57) S. Raju, Lessons from the Information Paradox, arXiv 2012.05770 (2020)
  • (58) D. Harlow, TASI Lectures on the Emergence of Bulk Physics in AdS/CFT, PoS TASI2017, 002 (2018). DOI 10.22323/1.305.0002
  • (59) A. Jahn, J. Eisert, Holographic tensor network models and quantum error correction: a topical review, Quantum Sci. Technol. 6(3), 033002 (2021). DOI 10.1088/2058-9565/ac0293
  • (60) B. Chen, B. Czech, Z.z. Wang, Quantum information in holographic duality, Rept. Prog. Phys. 85(4), 046001 (2022). DOI 10.1088/1361-6633/ac51b5
  • (61) A. Almheiri, T. Hartman, J. Maldacena, E. Shaghoulian, A. Tajdini, The entropy of Hawking radiation, Rev. Mod. Phys. 93(3), 035002 (2021). DOI 10.1103/RevModPhys.93.035002
  • (62) R. Mahajan, Recent Progress on the Black Hole Information Paradox, Reson. 26, 33 (2021). DOI 10.1007/s12045-020-1103-y
  • (63) A. Lewkowycz, J. Maldacena, Generalized gravitational entropy, Journal of High Energy Physics 2013(8), 90 (2013). DOI 10.1007/JHEP08(2013)090. URL https://doi.org/10.1007/JHEP08(2013)090
  • (64) X. Dong, A. Lewkowycz, M. Rangamani, Deriving covariant holographic entanglement, Journal of High Energy Physics 2016(11) (2016). DOI 10.1007/jhep11(2016)028. URL http://dx.doi.org/10.1007/JHEP11(2016)028
  • (65) M. Headrick, T. Takayanagi, A Holographic proof of the strong subadditivity of entanglement entropy, Phys. Rev. D 76, 106013 (2007). DOI 10.1103/PhysRevD.76.106013
  • (66) P. Hayden, M. Headrick, A. Maloney, Holographic Mutual Information is Monogamous, Phys. Rev. D 87(4), 046003 (2013). DOI 10.1103/PhysRevD.87.046003
  • (67) N. Bao, S. Nezami, H. Ooguri, B. Stoica, J. Sully, M. Walter, The Holographic Entropy Cone, JHEP 09, 130 (2015). DOI 10.1007/JHEP09(2015)130
  • (68) V.E. Hubeny, M. Rangamani, M. Rota, The holographic entropy arrangement, Fortsch. Phys. 67(4), 1900011 (2019). DOI 10.1002/prop.201900011
  • (69) V.E. Hubeny, M. Rangamani, M. Rota, Holographic entropy relations, Fortsch. Phys. 66(11-12), 1800067 (2018). DOI 10.1002/prop.201800067
  • (70) M. Srednicki, Entropy and area, Phys. Rev. Lett. 71, 666 (1993). DOI 10.1103/PhysRevLett.71.666
  • (71) L. Susskind, J. Uglum, Black hole entropy in canonical quantum gravity and superstring theory, Phys. Rev. D 50, 2700 (1994). DOI 10.1103/PhysRevD.50.2700
  • (72) D.N. Kabat, Black hole entropy and entropy of entanglement, Nucl. Phys. B 453, 281 (1995). DOI 10.1016/0550-3213(95)00443-V
  • (73) F. Larsen, F. Wilczek, Renormalization of black hole entropy and of the gravitational coupling constant, Nucl. Phys. B 458, 249 (1996). DOI 10.1016/0550-3213(95)00548-X
  • (74) T. Jacobson, Black hole entropy and induced gravity, arXiv gr-qc/9404039 (1994)
  • (75) A. Belin, N. Iqbal, S.F. Lokhande, Bulk entanglement entropy in perturbative excited states, SciPost Phys. 5(3), 024 (2018). DOI 10.21468/SciPostPhys.5.3.024
  • (76) A. Belin, N. Iqbal, J. Kruthoff, Bulk entanglement entropy for photons and gravitons in AdS3, SciPost Phys. 8(5), 075 (2020). DOI 10.21468/SciPostPhys.8.5.075
  • (77) A. Belin, S. Colin-Ellerin, Bootstrapping Quantum Extremal Surfaces I: The Area Operator, arXiv 2107.07516 (2021)
  • (78) L.A. Rubel, Necessary and sufficient conditions for carlson’s theorem on entire functions, Transactions of the American Mathematical Society 83(2), 417 (1956). URL http://www.jstor.org/stable/1992882
  • (79) R. Bousso, B. Freivogel, S. Leichenauer, V. Rosenhaus, C. Zukowski, Null Geodesics, Local CFT Operators and AdS/CFT for Subregions, Phys. Rev. D 88, 064057 (2013). DOI 10.1103/PhysRevD.88.064057
  • (80) B. Czech, J.L. Karczmarek, F. Nogueira, M. Van Raamsdonk, The Gravity Dual of a Density Matrix, Class. Quant. Grav. 29, 155009 (2012). DOI 10.1088/0264-9381/29/15/155009
  • (81) R. Bousso, S. Leichenauer, V. Rosenhaus, Light-sheets and AdS/CFT, Phys. Rev. D 86, 046009 (2012). DOI 10.1103/PhysRevD.86.046009
  • (82) V.E. Hubeny, M. Rangamani, Causal holographic information, Journal of High Energy Physics 2012(6), 114 (2012). DOI 10.1007/JHEP06(2012)114. URL https://doi.org/10.1007/JHEP06(2012)114
  • (83) A.C. Wall, The generalized second law implies a quantum singularity theorem, Classical and Quantum Gravity 30(16), 165003 (2013). DOI 10.1088/0264-9381/30/16/165003. URL https://doi.org/10.1088/0264-9381/30/16/165003
  • (84) E.H. Lieb, M.B. Ruskai, A fundamental property of quantum-mechanical entropy, Phys. Rev. Lett. 30, 434 (1973). DOI 10.1103/PhysRevLett.30.434. URL https://link.aps.org/doi/10.1103/PhysRevLett.30.434
  • (85) E.H. Lieb, M.B. Ruskai, Proof of the strong subadditivity of quantum-mechanical entropy, J. Math. Phys. 14, 1938 (1973). DOI 10.1063/1.1666274
  • (86) T. Nishioka, S. Ryu, T. Takayanagi, Holographic Entanglement Entropy: An Overview, J. Phys. A 42, 504008 (2009). DOI 10.1088/1751-8113/42/50/504008
  • (87) D. Marolf, A.C. Wall, Z. Wang, Restricted Maximin surfaces and HRT in generic black hole spacetimes, JHEP 05, 127 (2019). DOI 10.1007/JHEP05(2019)127
  • (88) C. Akers, N. Engelhardt, G. Penington, M. Usatyuk, Quantum Maximin Surfaces, JHEP 08, 140 (2020). DOI 10.1007/JHEP08(2020)140
  • (89) C. Akers, S. Hernández-Cuenca, P. Rath, Quantum Extremal Surfaces and the Holographic Entropy Cone, arXiv 2108.07280 (2021)
  • (90) M. Henningson, K. Skenderis, The Holographic Weyl anomaly, JHEP 07, 023 (1998). DOI 10.1088/1126-6708/1998/07/023
  • (91) V. Balasubramanian, P. Kraus, A stress tensor for anti-de sitter gravity, Communications in Mathematical Physics 208(2), 413 (1999). DOI 10.1007/s002200050764. URL http://dx.doi.org/10.1007/s002200050764
  • (92) S. de Haro, S.N. Solodukhin, K. Skenderis, Holographic reconstruction of space-time and renormalization in the AdS / CFT correspondence, Commun. Math. Phys. 217, 595 (2001). DOI 10.1007/s002200100381
  • (93) K. Skenderis, Lecture notes on holographic renormalization, Classical and Quantum Gravity 19(22), 5849 (2002). DOI 10.1088/0264-9381/19/22/306. URL http://dx.doi.org/10.1088/0264-9381/19/22/306
  • (94) I. Kanitscheider, K. Skenderis, M. Taylor, Precision holography for non-conformal branes, JHEP 09, 094 (2008). DOI 10.1088/1126-6708/2008/09/094
  • (95) D.T. Son, A.O. Starinets, Minkowski space correlators in AdS / CFT correspondence: Recipe and applications, JHEP 09, 042 (2002). DOI 10.1088/1126-6708/2002/09/042
  • (96) K. Skenderis, B.C. van Rees, Real-time gauge/gravity duality: Prescription, Renormalization and Examples, JHEP 05, 085 (2009). DOI 10.1088/1126-6708/2009/05/085
  • (97) C.P. Herzog, D.T. Son, Schwinger-Keldysh propagators from AdS/CFT correspondence, JHEP 03, 046 (2003). DOI 10.1088/1126-6708/2003/03/046
  • (98) P. Glorioso, M. Crossley, H. Liu, A prescription for holographic Schwinger-Keldysh contour in non-equilibrium systems, arXiv 1812.08785 (2018)
  • (99) L. Susskind, E. Witten, The Holographic bound in anti-de Sitter space, arXiv hep-th/9805114 (1998)
  • (100) T. Banks, M.R. Douglas, G.T. Horowitz, E.J. Martinec, AdS dynamics from conformal field theory, arXiv hep-th/9808016 (1998)
  • (101) H. Lehmann, K. Symanzik, W. Zimmermann, Zur Formulierung quantisierter Feldtheorien, Nuovo Cimento Serie 1(1), 205 (1955). DOI 10.1007/BF02731765
  • (102) S.B. Giddings, The Boundary S matrix and the AdS to CFT dictionary, Phys. Rev. Lett. 83, 2707 (1999). DOI 10.1103/PhysRevLett.83.2707
  • (103) D. Harlow, D. Stanford, Operator Dictionaries and Wave Functions in AdS/CFT and dS/CFT, arXiv 1104.2621 (2011)
  • (104) A. Hamilton, D. Kabat, G. Lifschytz, D.A. Lowe, Holographic representation of local bulk operators, Phys. Rev. D 74, 066009 (2006). DOI 10.1103/PhysRevD.74.066009. URL https://link.aps.org/doi/10.1103/PhysRevD.74.066009
  • (105) A. Hamilton, D. Kabat, G. Lifschytz, D.A. Lowe, Local bulk operators in ads/cft correspondence: A holographic description of the black hole interior, Phys. Rev. D 75, 106001 (2007). DOI 10.1103/PhysRevD.75.106001. URL https://link.aps.org/doi/10.1103/PhysRevD.75.106001
  • (106) R. Haag, Local quantum physics: Fields, particles, algebras (1992)
  • (107) J.J. Bisognano, E.H. Wichmann, On the Duality Condition for Quantum Fields, J. Math. Phys. 17, 303 (1976). DOI 10.1063/1.522898
  • (108) P.D. Hislop, R. Longo, Modular structure of the local algebras associated with the free massless scalar field theory, Communications in Mathematical Physics 84(1), 71 (1982). DOI 10.1007/BF01208372. URL https://doi.org/10.1007/BF01208372
  • (109) H. Casini, M. Huerta, R.C. Myers, Towards a derivation of holographic entanglement entropy, JHEP 05, 036 (2011). DOI 10.1007/JHEP05(2011)036
  • (110) F. Hiai, D. Petz, The proper formula for relative entropy and its asymptotics in quantum probability, Comm. Math. Phys. 143, 99 (1991). DOI 10.1007/BF02100287
  • (111) E. Witten, Aps medal for exceptional achievement in research: Invited article on entanglement properties of quantum field theory, Rev. Mod. Phys. 90, 045003 (2018). DOI 10.1103/RevModPhys.90.045003. URL https://link.aps.org/doi/10.1103/RevModPhys.90.045003
  • (112) M.A. Nielsen, I.L. Chuang, Quantum Computation and Quantum Information (Cambridge University Press, Cambridge, 2000)
  • (113) A. Uhlmann, Relative entropy and the wigner-yanase-dyson-lieb concavity in an interpolation theory, Communications in Mathematical Physics 54(1), 21 (1977)
  • (114) D.D. Blanco, H. Casini, L.Y. Hung, R.C. Myers, Relative Entropy and Holography, JHEP 08, 060 (2013). DOI 10.1007/JHEP08(2013)060
  • (115) D.L. Jafferis, S.J. Suh, The Gravity Duals of Modular Hamiltonians, JHEP 09, 068 (2016). DOI 10.1007/JHEP09(2016)068
  • (116) D. PETZ, C. GHINEA, Introduction to quantum fisher information, Quantum Probability and Related Topics (2011). DOI 10.1142/9789814338745˙0015. URL http://dx.doi.org/10.1142/9789814338745_0015
  • (117) N. Lashkari, M.B. McDermott, M. Van Raamsdonk, Gravitational dynamics from entanglement “thermodynamics”, Journal of High Energy Physics 2014(4) (2014). DOI 10.1007/jhep04(2014)195. URL http://dx.doi.org/10.1007/JHEP04(2014)195
  • (118) T. Faulkner, M. Guica, T. Hartman, R.C. Myers, M. Van Raamsdonk, Gravitation from entanglement in holographic cfts, Journal of High Energy Physics 2014(3) (2014). DOI 10.1007/jhep03(2014)051. URL http://dx.doi.org/10.1007/JHEP03(2014)051
  • (119) S. Hollands, R.M. Wald, Stability of black holes and black branes, Communications in Mathematical Physics 321(3), 629 (2012). DOI 10.1007/s00220-012-1638-1. URL http://dx.doi.org/10.1007/s00220-012-1638-1
  • (120) V. Iyer, R.M. Wald, Some properties of Noether charge and a proposal for dynamical black hole entropy, Phys. Rev. D 50, 846 (1994). DOI 10.1103/PhysRevD.50.846
  • (121) N. Lashkari, M. Van Raamsdonk, Canonical energy is quantum fisher information, Journal of High Energy Physics 2016(4), 1 (2016). DOI 10.1007/jhep04(2016)153. URL http://dx.doi.org/10.1007/JHEP04(2016)153
  • (122) B. Czech, L. Lamprou, S. McCandlish, B. Mosk, J. Sully, A Stereoscopic Look into the Bulk, JHEP 07, 129 (2016). DOI 10.1007/JHEP07(2016)129
  • (123) J. de Boer, F.M. Haehl, M.P. Heller, R.C. Myers, Entanglement, holography and causal diamonds, JHEP 08, 162 (2016). DOI 10.1007/JHEP08(2016)162
  • (124) B. Carneiro da Cunha, M. Guica, Exploring the BTZ bulk with boundary conformal blocks, arXiv 1604.07383 (2016)
  • (125) S. Banerjee, J.W. Bryan, K. Papadodimas, S. Raju, A toy model of black hole complementarity, JHEP 05, 004 (2016). DOI 10.1007/JHEP05(2016)004
  • (126) B. Czech, L. Lamprou, S. Mccandlish, J. Sully, Modular Berry Connection for Entangled Subregions in AdS/CFT, Phys. Rev. Lett. 120(9), 091601 (2018). DOI 10.1103/PhysRevLett.120.091601
  • (127) B. Czech, J. De Boer, D. Ge, L. Lamprou, A modular sewing kit for entanglement wedges, JHEP 11, 094 (2019). DOI 10.1007/JHEP11(2019)094
  • (128) T. Faulkner, R.G. Leigh, O. Parrikar, H. Wang, Modular Hamiltonians for Deformed Half-Spaces and the Averaged Null Energy Condition, JHEP 09, 038 (2016). DOI 10.1007/JHEP09(2016)038
  • (129) H. Casini, E. Teste, G. Torroba, Modular Hamiltonians on the null plane and the Markov property of the vacuum state, J. Phys. A 50(36), 364001 (2017). DOI 10.1088/1751-8121/aa7eaa
  • (130) I. Klich, D. Vaman, G. Wong, Entanglement Hamiltonians for chiral fermions with zero modes, Phys. Rev. Lett. 119(12), 120401 (2017). DOI 10.1103/PhysRevLett.119.120401
  • (131) N. Lashkari, Modular Hamiltonian for Excited States in Conformal Field Theory, Phys. Rev. Lett. 117(4), 041601 (2016). DOI 10.1103/PhysRevLett.117.041601
  • (132) I. Klich, D. Vaman, G. Wong, Entanglement Hamiltonians and entropy in 1+1D chiral fermion systems, Phys. Rev. B 98, 035134 (2018). DOI 10.1103/PhysRevB.98.035134
  • (133) G. Sárosi, T. Ugajin, Modular hamiltonians of excited states, ope blocks and emergent bulk fields, Journal of High Energy Physics 2018(1) (2018). DOI 10.1007/jhep01(2018)012. URL http://dx.doi.org/10.1007/JHEP01(2018)012
  • (134) M. Junge, R. Renner, D. Sutter, M.M. Wilde, A. Winter, Universal recovery maps and approximate sufficiency of quantum relative entropy, Annales Henri Poincaré 19(10), 2955 (2018). DOI 10.1007/s00023-018-0716-0. URL http://dx.doi.org/10.1007/s00023-018-0716-0
  • (135) D. Kabat, G. Lifschytz, P. Nguyen, D. Sarkar, Endpoint contributions to excited-state modular Hamiltonians, JHEP 12, 128 (2020). DOI 10.1007/JHEP12(2020)128
  • (136) D. Kabat, G. Lifschytz, P. Nguyen, D. Sarkar, Light-ray moments as endpoint contributions to modular Hamiltonians, JHEP 09, 074 (2021). DOI 10.1007/JHEP09(2021)074
  • (137) T.G. Mertens, Towards black hole evaporation in jackiw-teitelboim gravity, Journal of High Energy Physics 2019(7) (2019). DOI 10.1007/jhep07(2019)097. URL http://dx.doi.org/10.1007/JHEP07(2019)097
  • (138) M. Rozali, J. Sully, M. Van Raamsdonk, C. Waddell, D. Wakeham, Information radiation in bcft models of black holes, Journal of High Energy Physics 2020(5) (2020). DOI 10.1007/jhep05(2020)004. URL http://dx.doi.org/10.1007/JHEP05(2020)004
  • (139) R. Jackiw, Lower dimensional gravity, Nuclear Physics B 252, 343 (1985). DOI https://doi.org/10.1016/0550-3213(85)90448-1. URL https://www.sciencedirect.com/science/article/pii/0550321385904481
  • (140) C. Teitelboim, Gravitation and hamiltonian structure in two spacetime dimensions, Physics Letters B 126(1), 41 (1983). DOI https://doi.org/10.1016/0370-2693(83)90012-6. URL https://www.sciencedirect.com/science/article/pii/0370269383900126
  • (141) A. Almheiri, J. Polchinski, Models of AdS2 backreaction and holography, JHEP 11, 014 (2015). DOI 10.1007/JHEP11(2015)014
  • (142) P. Calabrese, J. Cardy, Entanglement entropy and conformal field theory, J. Phys. A 42, 504005 (2009). DOI 10.1088/1751-8113/42/50/504005
  • (143) L. Randall, R. Sundrum, An alternative to compactification, Physical Review Letters 83(23), 4690 (1999). DOI 10.1103/physrevlett.83.4690. URL http://dx.doi.org/10.1103/PhysRevLett.83.4690
  • (144) A. Karch, L. Randall, Locally localized gravity, Journal of High Energy Physics 2001(05), 008 (2001). DOI 10.1088/1126-6708/2001/05/008. URL http://dx.doi.org/10.1088/1126-6708/2001/05/008
  • (145) J. Maldacena, L. Susskind, Cool horizons for entangled black holes, Fortschritte der Physik 61(9), 781 (2013). DOI 10.1002/prop.201300020. URL http://dx.doi.org/10.1002/prop.201300020
  • (146) P. Hayden, J. Preskill, Black holes as mirrors: quantum information in random subsystems, Journal of High Energy Physics 2007(09), 120 (2007). DOI 10.1088/1126-6708/2007/09/120. URL http://dx.doi.org/10.1088/1126-6708/2007/09/120
  • (147) A. Almheiri, R. Mahajan, J. Santos, Entanglement islands in higher dimensions, SciPost Physics 9(1) (2020). DOI 10.21468/scipostphys.9.1.001. URL http://dx.doi.org/10.21468/SciPostPhys.9.1.001
  • (148) H.Z. Chen, R.C. Myers, D. Neuenfeld, I.A. Reyes, J. Sandor, Quantum Extremal Islands Made Easy, Part I: Entanglement on the Brane, JHEP 10, 166 (2020). DOI 10.1007/JHEP10(2020)166
  • (149) H.Z. Chen, R.C. Myers, D. Neuenfeld, I.A. Reyes, J. Sandor, Quantum Extremal Islands Made Easy, Part II: Black Holes on the Brane, JHEP 12, 025 (2020). DOI 10.1007/JHEP12(2020)025
  • (150) A. Almheiri, R. Mahajan, J. Maldacena, Islands outside the horizon, arXiv 1910.11077 (2019)
  • (151) H.Z. Chen, Z. Fisher, J. Hernandez, R.C. Myers, S.M. Ruan, Information flow in black hole evaporation, Journal of High Energy Physics 2020(3) (2020). DOI 10.1007/jhep03(2020)152. URL http://dx.doi.org/10.1007/JHEP03(2020)152
  • (152) F.F. Gautason, L. Schneiderbauer, W. Sybesma, L. Thorlacius, Page Curve for an Evaporating Black Hole, JHEP 05, 091 (2020). DOI 10.1007/JHEP05(2020)091
  • (153) C. Akers, N. Engelhardt, D. Harlow, Simple holographic models of black hole evaporation, JHEP 08, 032 (2020). DOI 10.1007/JHEP08(2020)032
  • (154) T. Hartman, E. Shaghoulian, A. Strominger, Islands in Asymptotically Flat 2D Gravity, JHEP 07, 022 (2020). DOI 10.1007/JHEP07(2020)022
  • (155) T.J. Hollowood, S.P. Kumar, Islands and Page Curves for Evaporating Black Holes in JT Gravity, JHEP 08, 094 (2020). DOI 10.1007/JHEP08(2020)094
  • (156) T. Anegawa, N. Iizuka, Notes on islands in asymptotically flat 2d dilaton black holes, JHEP 07, 036 (2020). DOI 10.1007/JHEP07(2020)036
  • (157) K. Hashimoto, N. Iizuka, Y. Matsuo, Islands in Schwarzschild black holes, JHEP 06, 085 (2020). DOI 10.1007/JHEP06(2020)085
  • (158) V. Balasubramanian, A. Kar, O. Parrikar, G. Sárosi, T. Ugajin, Geometric secret sharing in a model of Hawking radiation, JHEP 01, 177 (2021). DOI 10.1007/JHEP01(2021)177
  • (159) M. Alishahiha, A. Faraji Astaneh, A. Naseh, Island in the presence of higher derivative terms, JHEP 02, 035 (2021). DOI 10.1007/JHEP02(2021)035
  • (160) H. Geng, A. Karch, Massive islands, JHEP 09, 121 (2020). DOI 10.1007/JHEP09(2020)121
  • (161) T. Hartman, Y. Jiang, E. Shaghoulian, Islands in cosmology, JHEP 11, 111 (2020). DOI 10.1007/JHEP11(2020)111
  • (162) Y. Chen, V. Gorbenko, J. Maldacena, Bra-ket wormholes in gravitationally prepared states, JHEP 02, 009 (2021). DOI 10.1007/JHEP02(2021)009
  • (163) C. Krishnan, Critical Islands, JHEP 01, 179 (2021). DOI 10.1007/JHEP01(2021)179
  • (164) M. Van Raamsdonk, Comments on wormholes, ensembles, and cosmology, arXiv 2008.02259 (2020)
  • (165) V. Balasubramanian, A. Kar, T. Ugajin, Islands in de Sitter space, JHEP 02, 072 (2021). DOI 10.1007/JHEP02(2021)072
  • (166) W. Sybesma, Pure de Sitter space and the island moving back in time, Class. Quant. Grav. 38(14), 145012 (2021). DOI 10.1088/1361-6382/abff9a
  • (167) H. Geng, Y. Nomura, H.Y. Sun, Information paradox and its resolution in de Sitter holography, Phys. Rev. D 103(12), 126004 (2021). DOI 10.1103/PhysRevD.103.126004
  • (168) H. Geng, S. Lüst, R.K. Mishra, D. Wakeham, Holographic BCFTs and Communicating Black Holes, jhep 08, 003 (2021). DOI 10.1007/JHEP08(2021)003
  • (169) D. Gottesman, Quantum error correction and fault-tolerance, Quantum Information Processing: From Theory to Experiment 199, 159 (2006)
  • (170) W.K. Wootters, W.H. Zurek, A single quantum cannot be cloned, Nature 299(5886), 802 (1982)
  • (171) E. Knill, R. Laflamme, Theory of quantum error-correcting codes, Physical Review A 55(2), 900 (1997)
  • (172) B.M. Terhal, Quantum error correction for quantum memories, Reviews of Modern Physics 87(2), 307 (2015)
  • (173) D. Kribs, R. Laflamme, D. Poulin, Unified and generalized approach to quantum error correction, Physical review letters 94(18), 180501 (2005)
  • (174) M.A. Nielsen, D. Poulin, Algebraic and information-theoretic conditions for operator quantum error correction, Physical Review A 75(6), 064304 (2007)
  • (175) C. Bény, A. Kempf, D.W. Kribs, Generalization of quantum error correction via the heisenberg picture, Physical review letters 98(10), 100502 (2007)
  • (176) F. Pastawski, B. Yoshida, D. Harlow, J. Preskill, Holographic quantum error-correcting codes: Toy models for the bulk/boundary correspondence, Journal of High Energy Physics 2015(6), 1 (2015)
  • (177) R. Cleve, D. Gottesman, H.K. Lo, How to share a quantum secret, Physical Review Letters 83(3), 648 (1999)
  • (178) F. Pastawski, J. Preskill, Code properties from holographic geometries, Physical Review X 7(2), 021022 (2017)
  • (179) E. Gesteau, M.J. Kang, Nonperturbative gravity corrections to bulk reconstruction, arXiv 2112.12789 (2021)
  • (180) P. Hayden, R. Jozsa, D. Petz, A. Winter, Structure of states which satisfy strong subadditivity of quantum entropy with equality, Communications in mathematical physics 246(2), 359 (2004)
  • (181) A. Bhattacharyya, Z.S. Gao, L.Y. Hung, S.N. Liu, Exploring the Tensor Networks/AdS Correspondence, JHEP 08, 086 (2016). DOI 10.1007/JHEP08(2016)086
  • (182) R. Laflamme, C. Miquel, J.P. Paz, W.H. Zurek, Perfect quantum error correcting code, Physical Review Letters 77(1), 198 (1996)
  • (183) P. Hayden, S. Nezami, X.L. Qi, N. Thomas, M. Walter, Z. Yang, Holographic duality from random tensor networks, Journal of High Energy Physics 2016(11), 1 (2016)
  • (184) D. Harlow, The ryu–takayanagi formula from quantum error correction, Communications in Mathematical Physics 354(3), 865 (2017)
  • (185) H.F. Jia, M. Rangamani, Petz reconstruction in random tensor networks, Journal of High Energy Physics 2020(10), 1 (2020)
  • (186) E. Gesteau, M.J. Kang, The infinite-dimensional HaPPY code: entanglement wedge reconstruction and dynamics, arXiv 2005.05971 (2020)
  • (187) C.F. Chen, G. Penington, G. Salton, Entanglement wedge reconstruction using the petz map, Journal of High Energy Physics 2020(1), 1 (2020)
  • (188) M. Ohya, D. Petz, Quantum entropy and its use (Springer Science & Business Media, 2004)
  • (189) H. Barnum, E. Knill, Reversing quantum dynamics with near-optimal quantum and classical fidelity, Journal of Mathematical Physics 43(5), 2097 (2002)
  • (190) H.K. Ng, P. Mandayam, Simple approach to approximate quantum error correction based on the transpose channel, Physical Review A 81(6), 062342 (2010)
  • (191) D. Petz, Monotonicity of quantum relative entropy revisited, Reviews in Mathematical Physics 15(01), 79 (2003)
  • (192) K. Furuya, N. Lashkari, S. Ouseph, Real-space renormalization, error correction and conditional expectations, arXiv preprint arXiv:2012.14001 (2020)
  • (193) M. Ghodrati, Entanglement wedge reconstruction and correlation measures in mixed states: Modular flows versus quantum recovery channels, Physical Review D 104(4), 046004 (2021)
  • (194) S.S. Lee, Holographic description of quantum field theory, Nucl. Phys. B 832, 567 (2010). DOI 10.1016/j.nuclphysb.2010.02.022
  • (195) S.S. Lee, Quantum Renormalization Group and Holography, JHEP 01, 076 (2014). DOI 10.1007/JHEP01(2014)076
  • (196) N. Behr, A. Mukhopadhyay, Holography as a highly efficient renormalization group flow. II. An explicit construction, Phys. Rev. D 94(2), 026002 (2016). DOI 10.1103/PhysRevD.94.026002
  • (197) N. Behr, S. Kuperstein, A. Mukhopadhyay, Holography as a highly efficient renormalization group flow. I. Rephrasing gravity, Phys. Rev. D 94(2), 026001 (2016). DOI 10.1103/PhysRevD.94.026001
  • (198) G. Mandal, P. Nayak, Revisiting AdS/CFT at a finite radial cut-off, JHEP 12, 125 (2016). DOI 10.1007/JHEP12(2016)125
  • (199) B. Sathiapalan, H. Sonoda, A Holographic form for Wilson’s RG, Nucl. Phys. B 924, 603 (2017). DOI 10.1016/j.nuclphysb.2017.09.018
  • (200) I. Heemskerk, J. Penedones, J. Polchinski, J. Sully, Holography from Conformal Field Theory, JHEP 10, 079 (2009). DOI 10.1088/1126-6708/2009/10/079
  • (201) I. Heemskerk, J. Polchinski, Holographic and Wilsonian Renormalization Groups, JHEP 06, 031 (2011). DOI 10.1007/JHEP06(2011)031
  • (202) S. Kuperstein, A. Mukhopadhyay, The unconditional RG flow of the relativistic holographic fluid, JHEP 11, 130 (2011). DOI 10.1007/JHEP11(2011)130
  • (203) S. Kuperstein, A. Mukhopadhyay, Spacetime emergence via holographic RG flow from incompressible Navier-Stokes at the horizon, JHEP 11, 086 (2013). DOI 10.1007/JHEP11(2013)086
  • (204) A. Mukhopadhyay, Understanding the holographic principle via RG flow, Int. J. Mod. Phys. A 31(34), 1630059 (2016). DOI 10.1142/S0217751X16300593
  • (205) M.P. Heller, R.A. Janik, P. Witaszczyk, Hydrodynamic Gradient Expansion in Gauge Theory Plasmas, Phys. Rev. Lett. 110(21), 211602 (2013). DOI 10.1103/PhysRevLett.110.211602
  • (206) S. Bhattacharyya, V.E. Hubeny, S. Minwalla, M. Rangamani, Nonlinear Fluid Dynamics from Gravity, JHEP 02, 045 (2008). DOI 10.1088/1126-6708/2008/02/045
  • (207) R. Baier, P. Romatschke, D.T. Son, A.O. Starinets, M.A. Stephanov, Relativistic viscous hydrodynamics, conformal invariance, and holography, JHEP 04, 100 (2008). DOI 10.1088/1126-6708/2008/04/100
  • (208) G. Policastro, D.T. Son, A.O. Starinets, The Shear viscosity of strongly coupled N=4 supersymmetric Yang-Mills plasma, Phys. Rev. Lett. 87, 081601 (2001). DOI 10.1103/PhysRevLett.87.081601
  • (209) A. Almheiri, D. Marolf, J. Polchinski, J. Sully, Black Holes: Complementarity or Firewalls?, JHEP 02, 062 (2013). DOI 10.1007/JHEP02(2013)062
  • (210) M. Koashi, A. Winter, Monogamy of quantum entanglement and other correlations, Phys. Rev. A 69, 022309 (2004). DOI 10.1103/PhysRevA.69.022309. URL https://link.aps.org/doi/10.1103/PhysRevA.69.022309
  • (211) S.L. Braunstein, S. Pirandola, K. Życzkowski, Better late than never: Information retrieval from black holes, Phys. Rev. Lett. 110, 101301 (2013). DOI 10.1103/PhysRevLett.110.101301. URL https://link.aps.org/doi/10.1103/PhysRevLett.110.101301
  • (212) A. Kundu, Wormholes & Holography: An Introduction, arXiv 2110.14958 (2021)
  • (213) R. Bousso, A. Shahbazi-Moghaddam, Island Finder and Entropy Bound, Phys. Rev. D 103(10), 106005 (2021). DOI 10.1103/PhysRevD.103.106005
  • (214) A. Manu, K. Narayan, P. Paul, Cosmological singularities, entanglement and quantum extremal surfaces, JHEP 04, 200 (2021). DOI 10.1007/JHEP04(2021)200
  • (215) J. Kumar Basak, D. Basu, V. Malvimat, H. Parihar, G. Sengupta, Islands for Entanglement Negativity, arXiv 2012.03983 (2020)
  • (216) J. Kumar Basak, D. Basu, V. Malvimat, H. Parihar, G. Sengupta, Page Curve for Entanglement Negativity through Geometric Evaporation, arXiv 2106.12593 (2021)
  • (217) J. Kudler-Flam, S. Ryu, Entanglement negativity and minimal entanglement wedge cross sections in holographic theories, Phys. Rev. D 99(10), 106014 (2019). DOI 10.1103/PhysRevD.99.106014
  • (218) Y. Kusuki, J. Kudler-Flam, S. Ryu, Derivation of Holographic Negativity in AdS3/CFT2, Phys. Rev. Lett. 123(13), 131603 (2019). DOI 10.1103/PhysRevLett.123.131603
  • (219) H. Shapourian, S. Liu, J. Kudler-Flam, A. Vishwanath, Entanglement Negativity Spectrum of Random Mixed States: A Diagrammatic Approach, PRX Quantum 2(3), 030347 (2021). DOI 10.1103/PRXQuantum.2.030347
  • (220) J. Kudler-Flam, V. Narovlansky, S. Ryu, Negativity spectra in random tensor networks and holography, JHEP 02, 076 (2022). DOI 10.1007/JHEP02(2022)076
  • (221) X. Dong, S. McBride, W.W. Weng, Replica Wormholes and Holographic Entanglement Negativity, arXiv 2110.11947 (2021)
  • (222) E. Caceres, A. Kundu, A.K. Patra, S. Shashi, Page Curves and Bath Deformations, arXiv 2107.00022 (2021)
  • (223) Y. Chen, Pulling Out the Island with Modular Flow, JHEP 03, 033 (2020). DOI 10.1007/JHEP03(2020)033
  • (224) H. Liu, S. Vardhan, Entanglement entropies of equilibrated pure states in quantum many-body systems and gravity, PRX Quantum 2, 010344 (2021). DOI 10.1103/PRXQuantum.2.010344. URL https://link.aps.org/doi/10.1103/PRXQuantum.2.010344
  • (225) S. Vardhan, J. Kudler-Flam, H. Shapourian, H. Liu, Bound entanglement in thermalized states and black hole radiation, arXiv 2110.02959 (2021)
  • (226) P. Saad, S.H. Shenker, D. Stanford, JT gravity as a matrix integral, arXiv 1903.11115 (2019)
  • (227) D. Stanford, E. Witten, JT gravity and the ensembles of random matrix theory, Adv. Theor. Math. Phys. 24(6), 1475 (2020). DOI 10.4310/ATMP.2020.v24.n6.a4
  • (228) S.B. Giddings, A. Strominger, Baby Universes, Third Quantization and the Cosmological Constant, Nucl. Phys. B 321, 481 (1989). DOI 10.1016/0550-3213(89)90353-2
  • (229) J. Polchinski, A. Strominger, A Possible resolution of the black hole information puzzle, Phys. Rev. D 50, 7403 (1994). DOI 10.1103/PhysRevD.50.7403
  • (230) K. Papadodimas, S. Raju, An Infalling Observer in AdS/CFT, JHEP 10, 212 (2013). DOI 10.1007/JHEP10(2013)212
  • (231) K. Papadodimas, S. Raju, State-Dependent Bulk-Boundary Maps and Black Hole Complementarity, Phys. Rev. D 89(8), 086010 (2014). DOI 10.1103/PhysRevD.89.086010
  • (232) D. Harlow, Aspects of the Papadodimas-Raju Proposal for the Black Hole Interior, JHEP 11, 055 (2014). DOI 10.1007/JHEP11(2014)055
  • (233) P. Hayden, G. Penington, Learning the Alpha-bits of Black Holes, JHEP 12, 007 (2019). DOI 10.1007/JHEP12(2019)007
  • (234) C. Akers, S. Leichenauer, A. Levine, Large Breakdowns of Entanglement Wedge Reconstruction, Phys. Rev. D 100(12), 126006 (2019). DOI 10.1103/PhysRevD.100.126006
  • (235) G. Penington, Entanglement Wedge Reconstruction and the Information Paradox, JHEP 09, 002 (2020). DOI 10.1007/JHEP09(2020)002
  • (236) A.R. Brown, H. Gharibyan, G. Penington, L. Susskind, The Python’s Lunch: geometric obstructions to decoding Hawking radiation, JHEP 08, 121 (2020). DOI 10.1007/JHEP08(2020)121
  • (237) N. Engelhardt, G. Penington, A. Shahbazi-Moghaddam, Finding Pythons in Unexpected Places, arXiv 2105.09316 (2021)
  • (238) S.D. Mathur, The Fuzzball proposal for black holes: An Elementary review, Fortsch. Phys. 53, 793 (2005). DOI 10.1002/prop.200410203
  • (239) P. Hayden, G. Penington, Approximate Quantum Error Correction Revisited: Introducing the Alpha-Bit, Commun. Math. Phys. 374(2), 369 (2020). DOI 10.1007/s00220-020-03689-1
  • (240) C. Akers, G. Penington, Leading order corrections to the quantum extremal surface prescription, JHEP 04, 062 (2021). DOI 10.1007/JHEP04(2021)062
  • (241) L. Susskind, Computational Complexity and Black Hole Horizons, Fortsch. Phys. 64, 24 (2016). DOI 10.1002/prop.201500092. [Addendum: Fortsch.Phys. 64, 44–48 (2016)]
  • (242) A.R. Brown, D.A. Roberts, L. Susskind, B. Swingle, Y. Zhao, Holographic Complexity Equals Bulk Action?, Phys. Rev. Lett. 116(19), 191301 (2016). DOI 10.1103/PhysRevLett.116.191301
  • (243) A.R. Brown, D.A. Roberts, L. Susskind, B. Swingle, Y. Zhao, Complexity, action, and black holes, Phys. Rev. D 93(8), 086006 (2016). DOI 10.1103/PhysRevD.93.086006
  • (244) S. Chapman, G. Policastro, Quantum computational complexity from quantum information to black holes and back, Eur. Phys. J. C 82(2), 128 (2022). DOI 10.1140/epjc/s10052-022-10037-1
  • (245) B. Yoshida, A. Kitaev, Efficient decoding for the Hayden-Preskill protocol, arXiv 1710.03363 (2017)
  • (246) Y. Zhao, Petz map and Python’s lunch, JHEP 11, 038 (2020). DOI 10.1007/JHEP11(2020)038
  • (247) N. Engelhardt, G. Penington, A. Shahbazi-Moghaddam, A World without Pythons would be so Simple, arXiv 2102.07774 (2021)
  • (248) A. Levine, A. Shahbazi-Moghaddam, R.M. Soni, Seeing the entanglement wedge, JHEP 06, 134 (2021). DOI 10.1007/JHEP06(2021)134
  • (249) D.L. Jafferis, L. Lamprou, Inside the hologram: reconstructing the bulk observer’s experience, JHEP 03, 084 (2022). DOI 10.1007/JHEP03(2022)084
  • (250) S. Leutheusser, H. Liu, Emergent times in holographic duality, arXiv 2112.12156 (2021)
  • (251) Y. Sekino, L. Susskind, Fast Scramblers, JHEP 10, 065 (2008). DOI 10.1088/1126-6708/2008/10/065
  • (252) S.H. Shenker, D. Stanford, Black holes and the butterfly effect, JHEP 03, 067 (2014). DOI 10.1007/JHEP03(2014)067
  • (253) A. Bhattacharyya, L.K. Joshi, B. Sundar, Quantum Information Scrambling: From Holography to Quantum Simulators, arXiv 2111.11945 (2021)
  • (254) L. Susskind, L. Thorlacius, J. Uglum, The stretched horizon and black hole complementarity, Phys. Rev. D 48, 3743 (1993). DOI 10.1103/PhysRevD.48.3743. URL https://link.aps.org/doi/10.1103/PhysRevD.48.3743
  • (255) L. Susskind, J. Uglum, Black hole entropy in canonical quantum gravity and superstring theory, Phys. Rev. D 50, 2700 (1994). DOI 10.1103/PhysRevD.50.2700. URL https://link.aps.org/doi/10.1103/PhysRevD.50.2700
  • (256) D. Harlow, P. Hayden, Quantum Computation vs. Firewalls, JHEP 06, 085 (2013). DOI 10.1007/JHEP06(2013)085
  • (257) S. Aaronson, (2016), vol. 1607.05256
  • (258) I. Kim, E. Tang, J. Preskill, The ghost in the radiation: Robust encodings of the black hole interior, JHEP 06, 031 (2020). DOI 10.1007/JHEP06(2020)031
  • (259) T. Banks, W. Fischler, S.H. Shenker, L. Susskind, M theory as a matrix model: A Conjecture, Phys. Rev. D 55, 5112 (1997). DOI 10.1103/PhysRevD.55.5112
  • (260) T. Kibe, A. Mukhopadhyay, H. Swain, A. Soloviev, sl(2,r)𝑠𝑙2𝑟sl(2,r) lattices as information processors, Phys. Rev. D 102, 086008 (2020). DOI 10.1103/PhysRevD.102.086008. URL https://link.aps.org/doi/10.1103/PhysRevD.102.086008
  • (261) S.W. Hawking, M.J. Perry, A. Strominger, Soft hair on black holes, Phys. Rev. Lett. 116, 231301 (2016). DOI 10.1103/PhysRevLett.116.231301. URL https://link.aps.org/doi/10.1103/PhysRevLett.116.231301
  • (262) A. Strominger, Black Hole Information Revisited (2020), vol. 1706.07143. DOI 10.1142/9789811203961˙0010
  • (263) J.M. Maldacena, J. Michelson, A. Strominger, Anti-de Sitter fragmentation, JHEP 02, 011 (1999). DOI 10.1088/1126-6708/1999/02/011
  • (264) S. Sachdev, J. Ye, Gapless spin-fluid ground state in a random quantum heisenberg magnet, Physical Review Letters 70(21), 3339 (1993). DOI 10.1103/physrevlett.70.3339. URL http://dx.doi.org/10.1103/PhysRevLett.70.3339
  • (265) A. Kitaev. A simple model of quantum holography. https://online.kitp.ucsb.edu/online/entangled15/kitaev/ (Talks at KITP, April 7, 2015 and May 27, 2015)
  • (266) L.K. Joshi, A. Mukhopadhyay, A. Soloviev, Time-dependent nAdS2𝑛𝐴𝑑subscript𝑆2n{AdS}_{2} holography with applications, Phys. Rev. D 101, 066001 (2020). DOI 10.1103/PhysRevD.101.066001. URL https://link.aps.org/doi/10.1103/PhysRevD.101.066001
  • (267) E. Verlinde, H. Verlinde, Black Hole Entanglement and Quantum Error Correction, JHEP 10, 107 (2013). DOI 10.1007/JHEP10(2013)107
  • (268) S.D. Mathur, The Information paradox and the infall problem, Class. Quant. Grav. 28, 125010 (2011). DOI 10.1088/0264-9381/28/12/125010
  • (269) S.G. Avery, B.D. Chowdhury, A. Puhm, Unitarity and fuzzball complementarity: ’Alice fuzzes but may not even know it!’, JHEP 09, 012 (2013). DOI 10.1007/JHEP09(2013)012
  • (270) B. Swingle, Entanglement Renormalization and Holography, Phys. Rev. D 86, 065007 (2012). DOI 10.1103/PhysRevD.86.065007
  • (271) M. Miyaji, T. Takayanagi, Surface/State Correspondence as a Generalized Holography, PTEP 2015(7), 073B03 (2015). DOI 10.1093/ptep/ptv089
  • (272) P. Caputa, N. Kundu, M. Miyaji, T. Takayanagi, K. Watanabe, Anti-de Sitter Space from Optimization of Path Integrals in Conformal Field Theories, Phys. Rev. Lett. 119(7), 071602 (2017). DOI 10.1103/PhysRevLett.119.071602
  • (273) C. Akers, P. Rath, Holographic Renyi Entropy from Quantum Error Correction, JHEP 05, 052 (2019). DOI 10.1007/JHEP05(2019)052
  • (274) J. Goold, M. Huber, A. Riera, L.d. Rio, P. Skrzypczyk, The role of quantum information in thermodynamics—a topical review, Journal of Physics A: Mathematical and Theoretical 49(14), 143001 (2016). DOI 10.1088/1751-8113/49/14/143001. URL http://dx.doi.org/10.1088/1751-8113/49/14/143001
  • (275) N. Yunger Halpern, J.M. Renes, Beyond heat baths: Generalized resource theories for small-scale thermodynamics, Phys. Rev. E 93, 022126 (2016). DOI 10.1103/PhysRevE.93.022126. URL https://link.aps.org/doi/10.1103/PhysRevE.93.022126
  • (276) N. Yunger Halpern, P. Faist, J. Oppenheim, A. Winter, Microcanonical and resource-theoretic derivations of the thermal state of a quantum system with noncommuting charges, Nature Communications 7(1) (2016). DOI 10.1038/ncomms12051. URL http://dx.doi.org/10.1038/ncomms12051
  • (277) Y. Guryanova, S. Popescu, A.J. Short, R. Silva, P. Skrzypczyk, Thermodynamics of quantum systems with multiple conserved quantities, Nature Communications 7(1) (2016). DOI 10.1038/ncomms12049. URL http://dx.doi.org/10.1038/ncomms12049
  • (278) E. Chitambar, G. Gour, Quantum resource theories, Rev. Mod. Phys. 91, 025001 (2019). DOI 10.1103/RevModPhys.91.025001. URL https://link.aps.org/doi/10.1103/RevModPhys.91.025001
  • (279) T. Kibe, A. Mukhopadhyay, P. Roy, Quantum Thermodynamics of Holographic Quenches and Bounds on the Growth of Entanglement from the Quantum Null Energy Condition, Phys. Rev. Lett. 128(19), 191602 (2022). DOI 10.1103/PhysRevLett.128.191602
  • (280) A. Banerjee, T. Kibe, N. Mittal, A. Mukhopadhyay, P. Roy, Erasure tolerant quantum memory and the quantum null energy condition in holographic systems, arXiv 2202.00022 (2022)
  • (281) H. Geng, A. Karch, C. Perez-Pardavila, S. Raju, L. Randall, M. Riojas, S. Shashi, Information Transfer with a Gravitating Bath, SciPost Phys. 10(5), 103 (2021). DOI 10.21468/SciPostPhys.10.5.103
  • (282) H. Geng, A. Karch, C. Perez-Pardavila, S. Raju, L. Randall, M. Riojas, S. Shashi, Inconsistency of islands in theories with long-range gravity, JHEP 01, 182 (2022). DOI 10.1007/JHEP01(2022)182
  • (283) E. Shaghoulian, The central dogma and cosmological horizons, arXiv 2110.13210 (2021)
  • (284) A. Bzowski, P. McFadden, K. Skenderis, Holography for inflation using conformal perturbation theory, JHEP 04, 047 (2013). DOI 10.1007/JHEP04(2013)047