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

111[email protected]

1]Yukawa Institute for Theoretical Physics, Kyoto university, Kitashirakawa Oiwakecho, Sakyo-ku, Kyoto, 606-8502, Japan 2]Photon Science Center, Graduate School of Engineering, The University of Tokyo, Bunkyo-ku, Tokyo 113-8656, Japan 3]JST, PRESTO, 4-1-8 Honcho, Kawaguchi, Saitama, 332-0012, Japan 4]Department of Communication Engineering and Informatics, Graduate School of Informatics and Engineering, The University of Electro-Communications, Tokyo 182-8585, Japan

Black holes as clouded mirrors: the Hayden-Preskill protocol with symmetry

Yoshifumi Nakata    Eyuri Wakakuwa    Masato Koashi [ [ [ [
Abstract

The Hayden-Preskill protocol is a qubit-toy model of the black hole information paradox. Based on the assumption of scrambling, it was revealed that quantum information is instantly leaked out from the quantum many-body system that models a black hole. In this paper, we extend the protocol to the case where the system has symmetry and investigate how the symmetry affects the leakage of information. We especially focus on the conservation of the number of up-spins. Developing a partial decoupling approach, we first show that the symmetry induces a delay of leakage and an information remnant. We then clarify the physics behind them: the delay is characterized by thermodynamic properties of the system associated with the symmetry, and the information remnant is closely related to the symmetry-breaking of the initial state. These relations bridge the information leakage problem to macroscopic physics of quantum many-body systems and allow us to investigate the information leakage only in terms of physical properties of the system.

1 Introduction

Black holes are the most peculiar objects in the universe. While macroscopic properties of black holes can be fairly understood by general relativity, finding microscopic descriptions of black holes has been a central problem in fundamental physics. A significant step was made by the discovery of Hawking radiation [1, 2]: due to a quantum effect, a quantum black hole emits thermal radiation. This discovery raises a question about whether the radiation carries away the information in the black hole. Although information leakage is unlikely in the classical case due to the no-hair theorem [3, 4, 5], the holographic principle indicates that information should leak out as a black hole evaporates. This conflict is known as the information paradox.

A novel approach to the information paradox was proposed from the theory of quantum information, known as the Hayden-Preskill (HP) protocol [6]. In the protocol, a quantum black hole is modeled by a quantum many-body system consisting of a large number of qubits. After scrambling dynamics of the system, which is typically the case in quantum chaotic systems, the system is split into two random subsystems, one is a remaining black hole and the other is the ‘Hawking radiation’. Due to the scrambling dynamics, it was shown that information stored in the original many-body system can be recovered from the ‘radiation’ even if the size of the corresponding subsystem is small. This phenomena is often expressed as that information is quickly leaked out from the black hole. This result spiked various research topics over quantum gravity, quantum chaos, and quantum information [7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29], including the trials of experimentally simulating quantum gravity in a lab [30, 31, 32].

In this paper, we study the HP protocol when the quantum many-body system has symmetry. Since quantum many-body systems typically have symmetry, it is important to clarify how symmetry affects the protocol. We especially consider the system with conservation of the number of up-spins. The conservation law makes it difficult to fully investigate the protocol by the technique used in [6]. In [14, 15], the difficulty was avoided by considering restricted cases with fixed conserved quantities, and possible connections to other conceptual puzzles of quantum black holes [33, 34, 35, 36] were argued. Here, we develop a quantum information-theoretic technique to simultaneously deal with quantum information and symmetry, and investigate the HP protocol in general cases.

We first show that the symmetry changes when and how the information leaks out from the black hole. More specifically, we show the existence of a delay of information leakage and an information remnant, which are numerically shown to be macroscopically large for certain conditions of the initial many-body system. The technique we have developed may be of independent interest in relation to covariant quantum error-correcting codes [37, 38, 39, 40]. Therein, a lower bound on the recovery error, which corresponds to the information remnant in our terminology, was intensely studied for general types of symmetry. We then clarify the physics behind these phenomena: the delay and the information remnant are, respectively, characterized by thermodynamic properties and symmetry-breaking of the many-body system. These relations unveil non-trivial connections between quantum information in a many-body system and the macroscopic physics associated with the symmetry. Although we show the connections for the systems conserving the number of up-spins, we expect that they hold for a wider class of symmetries as far as they are global and abelian, such as U(1)U(1) symmetry.

The paper is organized as follows. For clarity of the presentation, we provide a brief overview of our results in Sec. 2. We start the main analysis with introducing our notation in Sec. 3. In Sec. 4, we overview the original HP protocol, and propose a modified protocol with symmetry. The partial decoupling is explained and is applied to the HP protocol with symmetry in Sec. 5. The physics behind the delay and the information remnant is discussed in Sec. 6. We conclude this paper with discussions in Sec. 7. Technical statements are proven in Appendices.

2 Overview of our results

2.1 Brief introduction of the Hayden-Preskill protocol

The HP protocol is a quantum information-theoretic toy model of a quantum black hole. In the model, a quantum black hole is represented by a closed quantum many-body system that undergoes unitary time evolution. After some time, a qubit is displaced from the system, which models an emission of a single qubit as the Hawking radiation from the black hole. The rest of the system again undergoes unitary time evolution, and another qubit is displaced. By repeating the unitary time evolution and displacement of a single qubit, the initial system is eventually split into two subsystems: one is considered to be a remaining black hole and the other is a number of qubits representing the Hawking radiation.

One may regard this model as the AdS/CFT correspondence of a quantum black hole, which predicts that a quantum black hole has a quantum mechanical description without gravity. This paper, however, does not argue the origin of the model and simply follows the original proposal of the protocol [6]. By focusing on the minimal setting, we can investigate the effect of symmetry to the HP protocol from a purely information-theoretic perspective.

Following the original modelling, we refer to the initial many-body system as a BH and to the random subsystems as a radiation and a remaining BH. The size of the radiation increases as time passes, and the size of a remaining BH decreases. Eventually, the whole BH becomes radiation.

A brief description of the HP protocol is as follows [6]. The full details are given in Sec. 4. See also Fig. 1 therein.

  1. 1.

    Quantum information source AA of kk qubits is thrown into an initial BH BinB_{\rm in} composed of NN qubits, which has already emitted some radiation BradB_{\rm rad}. The quantum information in AA is kept track of by a reference system RR.

  2. 2.

    The new BH S:=ABinS:=AB_{\rm in} of N+kN+k qubits undergoes scrambling dynamics and gradually emits new radiation, i.e., a random subsystem SradS_{\rm rad} of SS. The number of qubits in SradS_{\rm rad} is denoted by \ell. The BH shrinks to the remaining BH SinS_{\rm in} of N+kN+k-\ell qubits. As time goes on, \ell increases.

  3. 3.

    The quantum information originally stored in AA is tried to be recovered from the radiation BradB_{\rm rad} and SradS_{\rm rad}.

It was shown in Ref. [6] that, if the size \ell of the new radiation SradS_{\rm rad} exceeds a certain threshold, which is roughly given by

>k+NH(Bin)2,\ell>k+\frac{N-H(B_{\rm in})}{2}, (1)

where H(Bin)H(B_{\rm in}) is the entropy of the initial BH BinB_{\rm in}, the remaining BH SinS_{\rm in} is decoupled from the reference RR. As a result, the error in recovering the information from the radiations BradB_{\rm rad} and SradS_{\rm rad} decreases exponentially quickly as \ell increases. The threshold depends on the properties of the initial BH BinB_{\rm in}, and can be the same order of the number kk of qubits in the quantum information source AA.

2.2 The HP protocol with symmetry and the main result

The main concern in this paper is to clarify what would occur if the system has symmetry. Since the dynamics cannot be fully scrambling in the presence of symmetry, information should be leaked out from the system in a different manner. We particularly consider conservation of the number of up-spins, or equivalently, the ZZ-component of the angular momentum (ZZ-axis AM for short). Since the corresponding symmetry is a ZZ-axial symmetry, we call the system the Kerr BH in analogy with actual black holes. Throughout the paper, we denote by LL the mean of the ZZ-axis AM in the initial Kerr BH BinB_{\rm in} and by δL\delta L its standard deviation.

Assuming that the dynamics of the system is scrambling that respects the axial symmetry, we first show that partial decoupling occurs in the Kerr BH, rather than the full decoupling as in the case of BH without any symmetry. This results in two substantial changes, the delay of information leakage and the information remnant. The delay means that, compared to the BH without any symmetry, one needs to collect more radiation in order to recover quantum information of AA. The information remnant means that a certain amount of the information of AA remains in the Kerr BH until it is all radiated. Depending on the ZZ-axis AM LL and its fluctuation δL\delta L of the initial Kerr BH BinB_{\rm in}, the delay ranges from O(k)O(\sqrt{k}) to macroscopically large values, and the information remnant varies from infinitesimal to constant.

Our result implies that, with certain initial conditions, the symmetry radically changes the behavior of how radiation carries information away from the BH. In particular, when the Kerr BHs has extremely large ZZ-axis AM or fluctuation, there is a possibility that the Kerr BHs do keep nearly all information of AA until the last moment. This is in sharp contrast to the HP protocol without symmetry that predicts an instant leakage of information. See Sec. 5 for details.

2.3 Physics behind the delay of information leakage and information remnant

To physically understand the delay and the information remnant, we further investigate them from different perspectives. The details are given in Sec. 6.

We argue that the delay is ascribed to the structure of the entanglement generated by the scrambling dynamics with symmetry. To capture the information-theoretic consequence of such entanglement, we introduce a new concept that we call clipping of entanglement, and propose the following condition for the information recovery from the radiation to be possible: for most of possible values of the ZZ-axis AM nn in the new radiation SradS_{\rm rad},

k<H(Bin)+log[dimnSraddimLnSin],k<H(B_{\rm in})+\log\biggl{[}\frac{\dim{\mathcal{H}}_{n}^{S_{\rm rad}}}{\dim{\mathcal{H}}_{L-n}^{S_{\rm in}}}\biggr{]}, (2)

where H(Bin)H(B_{\rm in}) is the entropy of the initial Kerr BH BinB_{\rm in}, nSrad{\mathcal{H}}_{n}^{S_{\rm rad}} is the subspace of the Hilbert space of SradS_{\rm rad} with the ZZ-axis AM nn, and LnSin{\mathcal{H}}^{S_{\rm in}}_{L-n} is that of the remaining Kerr BH SinS_{\rm in} with the ZZ-axis AM LnL-n. This condition indeed reproduces the results based on the partial decoupling. Moreover, the results of the BH without symmetry [6] can be also obtained: in this case, the condition simply reduces to

k<H(Bin)+log[dimSraddimSin],k<H(B_{\rm in})+\log\biggl{[}\frac{\dim{\mathcal{H}}^{S_{\rm rad}}}{\dim{\mathcal{H}}^{S_{\rm in}}}\biggr{]}, (3)

due to the absence of conserved quantities. From this, one obtains >k+(NH(Bin))/2\ell>k+(N-H(B_{\rm in}))/2, which is the same as Eq. (1).

A merit of the entanglement clipping is that the condition for the information leakage, Eq. (2), consists only of the entropy of the initial system and the dimension of the Hilbert spaces with respect to the conserved quantity. Hence, for quantum many-body systems with any extensive conserved quantity, we can easily compute the delay of information leakage.

Apart from the conciseness, the argument based on entanglement clipping has another merit that it unveils a non-trivial relation between the delay of information leakage and thermodynamic properties of the system. We first show that the delay is characterized by up to the second-order derivatives of the entropy in BinB_{\rm in} in terms of the ZZ-axis AM. The derivatives are then connected to thermodynamic properties of BinB_{\rm in} as follows. Let ω(T,λ)\omega(T,\lambda) be the state function of BinB_{\rm in} conjugate to the ZZ-axis AM, which equals to the angular velocity of the Kerr BH, and α(T,λ)\alpha(T,\lambda) be thermal sensitivity of the ZZ-axis AM. We can show that, when the initial Kerr BH BinB_{\rm in} has a small ZZ-axis AM,

delayLSkB|ω(T,λ)α(T,λ)|,{\rm delay}\propto\frac{L}{S}\sqrt{k_{B}|\omega(T,\lambda)\alpha(T,\lambda)|}, (4)

where kBk_{B} is the Boltzmann constant, and SS is the entropy of the initial Kerr BH. Note that both ω(T,λ)\omega(T,\lambda) and α(T,λ)\alpha(T,\lambda) depend on the temperature TT, but their product is temperature-independent. The formula for a large ZZ-axis AM is also obtained. This allows us to understand the delay in terms of thermodynamic properties of the initial system.

The information remnant can also be intuitively understood. The fact that partial decoupling occurs instead of full decoupling implies that the ZZ-axis AM of the remaining Kerr BH SinS_{\rm in} remains correlated to that with the reference RR. In other words, the information originally stored in AA can be partially obtained by measuring the ZZ-axis AM in SinS_{\rm in}. From the viewpoint of the radiation SradS_{\rm rad}, this implies that the information about the coherence of the ZZ-axis AM in AA cannot be fully accessed, leading to the information remnant. The amount of the information remnant is closely related to the fluctuation of the ZZ-axis AM in the remaining Kerr BH SinS_{\rm in}: if the fluctuation is small, the change in the ZZ-axis AM caused by throwing AA into the Kerr BH is easy to detect. We quantitatively confirm this expectation and show that

informationremnant\displaystyle{\rm information}\ {\rm remnant} 1δν2,\displaystyle\gtrsim\frac{1}{\sqrt{\langle\delta\nu^{2}\rangle}}, (5)

where δν2\sqrt{\langle\delta\nu^{2}\rangle} is the standard deviation of the ZZ-axis AM in the remaining Kerr BH SinS_{\rm in}. We further show that the fluctuation δν2\sqrt{\langle\delta\nu^{2}\rangle} is characterized by the degree ζ(Sin)\zeta(S_{\rm in}) of symmetry-breaking in SinS_{\rm in}, leading to

informationremnant\displaystyle{\rm information}\ {\rm remnant} 12ζ(Sin).\displaystyle\gtrsim\frac{1}{\sqrt{2\zeta(S_{\rm in})}}. (6)

This reveals that the information remnant is characterized by the degree of symmetry-breaking of the Kerr BH.

The relations, Eqs. (4) and (6), establish quantitative connections between the information leakage problem and macroscopic physics. We expect that they universally hold for any quantum chaotic systems with abelian symmetry. In those cases, the quantities related to the ZZ-axis AM in Eqs. (4) and (6) should be replaced by the corresponding conserved quantities.

A canonical instance is energy conservation, in which the relations similar to Eqs. (4) and (6) can be obtained. The delay of information leakage when energy is conserved should be given in terms of the heat capacity CVC_{V} of the quantum system as

delayESkB|CV|{\rm delay}\propto\frac{E}{S}\sqrt{k_{B}|C_{V}|} (7)

where EE and SS are the energy and the entropy of the initial BH, respectively, and the amount of information remnant is determined by the energy fluctuation of the BH.

2.4 Discussions

Our results reveal, both in quantitative and physically intuitive manners, the presence of symmetry leads to non-negligible changes in the process of information leakage in the HP protocol. The changes were quantified by using the method based on partial decoupling as in Subsec. 2.2, and were qualitatively estimated by physical quantities of the system as explained in Subsec. 2.3. We are mainly concerned with rotational symmetry in our analysis and show that both the delay of information leakage and the information remnant can be large for certain initial conditions. Our approach, methods, and results offer a solid basis for studying quantum information in the many-body systems with symmetry.

In the context of quantum gravity, one of our main contributions is to show that more careful treatments of the symmetry, including the energy conservation, will be needed to fully understand the information recovery from the Hawking radiation. In fact, the presence of symmetry can result in the situation where information thrown into a quantum black hole cannot be recovered until it is completely evaporated. This conclusion is in sharp contrast to that of the original analysis of the HP protocol without symmetry and tells us that taking the black hole’s symmetry, such as energy conservation, into account is crucial for understanding the quantum information perspective of quantum gravity through the HP protocol.

A couple of analyses were already carried out about the energy conservation in the HP protocol [14, 15]. Therein, the physical modes of the radiation responsible for the instant information recovery and the relations between the HP protocol and other puzzles in quantum gravity were discussed [33, 34, 35, 36]. Despite the fact that these studies shed novel light on quantum gravity, the arguments highly depend on information-theoretic assumptions about the details of the model, which are sometimes unimportant from the viewpoint of physics or even unphysical. Hence, the conclusions are strongly depending on information-theoretic details of the model: one may arrive at entirely different results depending of the assumptions. To avoid such undesired situations, understanding the information perspectives of quantum gravity without referring to too much details about quantum information will be helpful.

Our approach provides a sophisticated method to this end. The simple formulas obtained in this paper can be used to understand information recovery only in terms of physical quantities. For instance, Eq. (2) is a formula to investigate the information recovery only from the degeneracy of eigenspaces. This implies, in the case of energy conservation, that the information recovery can be assessed only from the energy spectrum of the Hamiltonian of a quantum black hole. If one accepts a controversial assumption that a quantum black hole has usual thermodynamic features, which may be justified by the AdS/CFT correspondence, Eq. (7) provides a more direct way of evaluating the information recovery only from macroscopic physical properties of the black hole. The information remnant induced by the energy conservation can be also discussed based on Eq. (5), which may however be negligible in the thermodynamic limit if a black hole has a typical amount of energy fluctuation.

To summarize, our analysis provides 1. a caution not to literally accept the instant recovery of quantum information from the Hawking radiation in a realistic situation, and 2. convenient tools of investigating the information recovery only in terms of the quantities often studied in physics. Although our result itself does not provide conclusive answers to the most intriguing questions, such as whether the process of information leakage is changed by energy conservation in the leading order, our approach will be a stepping stone toward the full understanding of the information recovery from the Hawking radiation in a realistic situation with energy conservation. To this end, a couple of physical properties of a quantum black hole, such as the energy spectrum, its initial state, and the radiation process with energy conservation, should be clarified.

3 Preliminaries

Throughout the paper, we denote the sequence {i,i+1,,j1,j}\{i,i+1,\dots,j-1,j\} of integers between two natural numbers ii and jj (iji\leq j) by [i,j][i,j]. The logarithm is always taken in base two.

The basic unit of quantum information is a qubit represented by a two-dimensional Hilbert space. The number of qubits in a system SS is denoted by |S||S|. We often write the relevant systems in the superscript, such as a Hilbert space S\mathcal{H}^{S} of a system SS, an operator XSRX^{SR} on SRSR, and a superoperator SB\mathcal{E}^{S\rightarrow B} from SS to BB. For superoperators from SS to itself, we denote it by S\mathcal{E}^{S}.

The partial trace over a system XX is denoted by TrX\operatorname{Tr}_{X}. A reduced operator on SS of ρSR\rho^{SR} is denoted simply by ρS\rho^{S}, that is, ρS=TrR[ρSR]\rho^{S}=\operatorname{Tr}_{R}[\rho^{SR}]. Furthermore, we denote (MSIR)ρSR(MSIR)(M^{S}\otimes I^{R})\rho^{SR}(M^{S\dagger}\otimes I^{R}), where II is the identity operator, by MSρSRMSM^{S}\rho^{SR}M^{S\dagger}. The identity superoperator on SS is denoted by idS{\rm id}^{S}.

3.1 Linear Operators and Superoperators

We denote a set of linear operators from A{\mathcal{H}}^{A} to B{\mathcal{H}}^{B} by (A,B)\mathcal{L}({\mathcal{H}}^{A},{\mathcal{H}}^{B}), and (A,A)\mathcal{L}({\mathcal{H}}^{A},{\mathcal{H}}^{A}) by (A)\mathcal{L}({\mathcal{H}}^{A}). We also use the following notation for the sets of positive semi-definite operators, quantum states, and sub-normalized states.

𝒫()={X():X0},\displaystyle\mathcal{P}(\mathcal{H})=\{X\in\mathcal{L}(\mathcal{H}):X\geq 0\}, (8)
𝒮()={ρ𝒫():Tr[ρ]=1},\displaystyle\mathcal{S}(\mathcal{H})=\{\rho\in\mathcal{P}(\mathcal{H}):\operatorname{Tr}[\rho]=1\}, (9)
𝒮()={ρ𝒫():Tr[ρ]1}.\displaystyle\mathcal{S}_{\leq}(\mathcal{H})=\{\rho\in\mathcal{P}(\mathcal{H}):\operatorname{Tr}[\rho]\leq 1\}. (10)

A maximally entangled state between SS and SS^{\prime}, where SS{\mathcal{H}}^{S}\cong{\mathcal{H}}^{S^{\prime}}, is denoted by ΦSS\Phi^{SS^{\prime}}. The completely mixed state on S{\mathcal{H}}^{S} is denoted by πS=IS/dS\pi^{S}=I^{S}/d_{S} (dS=dimSd_{S}={\rm dim}{\mathcal{H}}^{S}).

A purification of ρS𝒮(S)\rho^{S}\in{\mathcal{S}}({\mathcal{H}}^{S}) by RR (RS{\mathcal{H}}^{R}\cong{\mathcal{H}}^{S}) is denoted by |ρSR|\rho\rangle^{SR}, namely, TrR[|ρρ|SR]=ρS\operatorname{Tr}_{R}[|\rho\rangle\langle\rho|^{SR}]=\rho^{S}. For instance, since the marginal state of a maximally entangled state is the completely mixed state, we have ΦS=πS\Phi^{S}=\pi^{S}.

The fundamental superoperators are the conjugations by a unitary, an isometry, and a partial isometry. An isometry V𝒦V^{{\mathcal{H}}\rightarrow{\cal K}} is the linear operator such that

(V𝒦)V𝒦=I.(V^{{\mathcal{H}}\rightarrow{\cal K}})^{\dagger}V^{{\mathcal{H}}\rightarrow{\cal K}}=I^{{\mathcal{H}}}. (11)

When dim=dim𝒦\dim{\mathcal{H}}=\dim{\cal K}, the isometry is called a unitary and also satisfies V𝒦(V𝒦)=I𝒦V^{{\mathcal{H}}\rightarrow{\cal K}}(V^{{\mathcal{H}}\rightarrow{\cal K}})^{\dagger}=I^{{\cal K}}. A partial isometry is a linear operator from {\mathcal{H}} to 𝒦{\cal K} such that it is an isometry on its support. Projections, isometries, and unitaries are special classes of a partial isometry.

A quantum channel 𝒯SB\mathcal{T}^{S\rightarrow B} is a completely-positive (CP) and trace-preserving (TP) map. A map is called CP if (idS𝒯SB)(ρSS)0({\rm id}^{S^{\prime}}\otimes\mathcal{T}^{S\rightarrow B})(\rho^{S^{\prime}S})\geq 0 for any ρSS0\rho^{S^{\prime}S}\geq 0 and is TP if Tr[𝒯SB(ρS)]=Tr[ρS]\operatorname{Tr}[\mathcal{T}^{S\rightarrow B}(\rho^{S})]=\operatorname{Tr}[\rho^{S}]. We also say a superoperator 𝒯SB\mathcal{T}^{S\rightarrow B} is sub-unital and unital if 𝒯SB(IS)IB\mathcal{T}^{S\rightarrow B}(I^{S})\leq I^{B} and 𝒯SB(IS)=IB\mathcal{T}^{S\rightarrow B}(I^{S})=I^{B}, respectively.

For a quantum channel 𝒯SB{\cal T}^{S\rightarrow B}, a complementary channel is defined as follows. Let V𝒯SBEV_{{\cal T}}^{S\rightarrow BE} be a Steinspring dilation for a quantum channel 𝒯SB\mathcal{T}^{S\rightarrow B}, i.e., V𝒯SBEV^{S\rightarrow BE}_{{\cal T}} is an isometry satisfying

TrE[V𝒯SBEρS(V𝒯SBE)]=𝒯SB(ρS),\operatorname{Tr}_{E}[V_{{\cal T}}^{S\rightarrow BE}\rho^{S}(V_{{\cal T}}^{S\rightarrow BE})^{\dagger}]=\mathcal{T}^{S\rightarrow B}(\rho^{S}), (12)

for any ρS(S)\rho^{S}\in\mathcal{L}(\mathcal{H}^{S}). The map 𝒯¯SE\bar{\mathcal{T}}^{S\rightarrow E} defined by

𝒯¯SE(ρS):=TrB[V𝒯SBEρS(V𝒯SBE)],\bar{\mathcal{T}}^{S\rightarrow E}(\rho^{S}):=\operatorname{Tr}_{B}[V_{{\cal T}}^{S\rightarrow BE}\rho^{S}(V^{S\rightarrow BE}_{{\cal T}})^{\dagger}], (13)

for any ρS(S)\rho^{S}\in\mathcal{L}(\mathcal{H}^{S}) is called a complementary channel.

Any superoperator 𝒯SB\mathcal{T}^{S\rightarrow B} from SS to BB has a Choi-Jamiołkowski representation defined by

𝔍(𝒯SB):=(idS𝒯SB)(ΦSS).\mathfrak{J}(\mathcal{T}^{S\rightarrow B}):=({\rm id}^{S^{\prime}}\otimes\mathcal{T}^{S\rightarrow B})(\Phi^{SS^{\prime}}). (14)

This is sometimes called the channel-state duality.

3.2 Norm and Entropy

The Schatten pp-norm for a linear operator X(A,B)X\in\mathcal{L}({\mathcal{H}}^{A},{\mathcal{H}}^{B}) is defined by Xp:=(Tr[(XX)p/2])1/p\|X\|_{p}:=(\operatorname{Tr}[(XX^{\dagger})^{p/2}])^{1/p} (p[1,]p\in[1,\infty]). We particularly use the trace (p=1p=1) and the Hilbert-Schmidt (p=2p=2) norms. We define the fidelity between quantum states ρ\rho and σ\sigma by

F(ρ,σ):=ρσ12.F(\rho,\sigma):=\|\sqrt{\rho}\sqrt{\sigma}\|_{1}^{2}. (15)

The conditional min-entropy is defined by

Hmin(A|B)ρ=supσB𝒮(B)sup{λ|2λIAσBρAB0},H_{\rm min}(A|B)_{\rho}\\ =\sup_{\sigma^{B}\in\mathcal{S}(\mathcal{H}^{B})}\sup\{\lambda\in\mathbb{R}|2^{-\lambda}I^{A}\otimes\sigma^{B}-\rho^{AB}\geq 0\}, (16)

for ρAB𝒫()\rho^{AB}\in\mathcal{P}({\mathcal{H}}). This is a variant of the conditional entropy H(A|B)ρ:=H(AB)H(B)H(A|B)_{\rho}:=H(AB)-H(B), where H(A)=Tr[ρAlogρA]H(A)=-\operatorname{Tr}[\rho^{A}\log\rho^{A}] is the von Neumann entropy. When dimB=1{\rm dim}{\mathcal{H}}^{B}=1, Hmin(A|B)ρH_{\rm min}(A|B)_{\rho} reduces to the min-entropy

Hmin(A)ρ=sup{λ|2λIAρA0}.H_{\rm min}(A)_{\rho}=\sup\{\lambda\in\mathbb{R}|2^{-\lambda}I^{A}-\rho^{A}\geq 0\}. (17)

3.3 Haar scrambling

The Haar measure is often used to formulate the scrambling dynamics. The Haar measure 𝖧{\sf H} on a unitary group 𝖴(d){\sf U}(d) of finite degree dd is the unique left- and right- unitarily invariant probability measure. Namely, it satisfies,

for any 𝒲𝖴(d) and V𝖴(d),𝖧(V𝒲)=𝖧(𝒲V)=𝖧(𝒲).\text{for any $\mathcal{W}\subset{\sf U}(d)$ and $V\in{\sf U}(d)$},\\ {\sf H}(V\mathcal{W})={\sf H}(\mathcal{W}V)={\sf H}(\mathcal{W}). (18)

When a unitary UU is chosen with respect to the Haar measure 𝖧{\sf H}, denoted by U𝖧U\sim{\sf H}, it is called a Haar random unitary. Throughout the paper, the average of a function f(U)f(U) over a random unitary UμU\sim\mu, where μ\mu is a probability measure, is denoted by 𝔼Uμ[f(U)]\mathbb{E}_{U\sim\mu}[f(U)].

4 Hayden-Preskill protocol

We overview the general setting of the HP protocol in Subsec. 4.1 and review the results in Ref. [6] with a slight generalization in Subsec. 4.2. We then explain how symmetry shall change the situation in Subsec. 4.3. The diagram of the protocol is given in Fig. 1, and important quantities are summarized in Table 1.

Refer to caption
Figure 1: Diagram of the HP protocol. The blue lines represent trajectories of qubits, and the yellow wavy lines indicate that they may be entangled. The black region in the right-hand side represents the BH system, and the blue region in the left-hand side is a reference system that is virtually introduced so as to keep track of the quantum information of AA. The green boxes in the black region represent the internal scrambling dynamics of the BH. Each UN+kiU_{N+k-i} is internal unitary time-evolution acting on N+kiN+k-i qubits. The purple lines represent the radiation of the BH. The map 𝒟{\mathcal{D}} is the recovery operation applied by Bob, aiming at recovering the information originally stored in AA from the past radiation BradB_{\rm rad} and the new radiation SradS_{\rm rad}.
Table 1: Notation in the investigation of the BH information problem
BinB_{\rm in} (NN qubits) Initial BH.
BradB_{\rm rad} (NN^{\prime} qubits) “Past” radiation that purifies the initial BH BinB_{\rm in}.
AA (kk qubits) Quantum information source Alice throws into the BH BinB_{\rm in}.
RR (kk qubits) Reference system that is maximally entangled with AA.
SS (N+kN+k qubits) The BH after the information source AA is thrown in (S=ABin=SinSradS=AB_{\rm in}=S_{\rm in}S_{\rm rad}).
SradS_{\rm rad} (\ell qubits) New radiation from the BH SS.
SinS_{\rm in} (N+kN+k-\ell qubits) Remaining BH after SradS_{\rm rad} is evaporated.

4.1 General setting of the HP protocol

In the HP protocol, we consider the situation depicted in Fig. 1. In the protocol, the BH BinB_{\rm in} initially consists of NN qubits. The radiation emitted earlier, which we call the past radiation, is denoted by a quantum system BradB_{\rm rad}.

We particularly consider two types of the BH. One is a “pure” BH, where BinB_{\rm in} is in a pure state |ξBin|\xi\rangle^{B_{\rm in}}, and the other is a “mixed” BH, where BinB_{\rm in} is in a mixed state ξBin\xi^{B_{\rm in}}. The former models the quantum black hole that does not emit radiation yet or the one whose radiation has been measured by someone. The latter models a sufficiently old quantum black hole that has already emitted a large amount of radiation. For a mixed BH, the system BinBradB_{\rm in}B_{\rm rad} is in a pure state |ξBinBrad|\xi\rangle^{B_{\rm in}B_{\rm rad}}, which is a purification of ξBin\xi^{B_{\rm in}}.

The HP protocol is the following thought-experiment. At time T1T_{1}, a person, Alice, throws a quantum information source AA of kk qubits into the BH BinB_{\rm in}. The quantum information source is defined by introducing a reference system RR, and by setting the initial state between AA and RR to a maximally entangled state ΦAR\Phi^{AR}. The BH, now a composite system S:=ABinS:=AB_{\rm in}, undergoes the internal unitary dynamics and emits new radiation. We model this process by repeatedly applying an internal unitary dynamics followed by a displacement of a qubit in SS to the exterior of the BH. Since each radiation, i.e., the displacement of a qubit, shrinks the BH by one qubit, the unitary after ii qubits are radiated acts on N+kiN+k-i qubits. In Fig. 1, the internal unitary dynamics acting on the N+kiN+k-i qubits in the BH is denoted by UN+kiU_{N+k-i}.

Suppose that the BH emits \ell qubits of radiation, i.e., |Srad|=|S_{\rm rad}|=\ell, by time T2T_{2}. We call SradS_{\rm rad} the new radiation. The remaining BH SinS_{\rm in} is composed of N+kN+k-\ell qubits. Since the dynamics from time T1T_{1} to T2T_{2} is quantum mechanical, the dynamics should be represented by a quantum channel. From the construction, the dynamics SSinSrad\mathcal{L}^{S\rightarrow S_{\rm in}S_{\rm rad}} from SS to SinSradS_{\rm in}S_{\rm rad} is simply a product of unitary dynamics,

SSinSrad=𝒰N+k+1𝒰N+k+2𝒰N+k,\mathcal{L}^{S\rightarrow S_{\rm in}S_{\rm rad}}=\mathcal{U}_{N+k-\ell+1}\circ\mathcal{U}_{N+k-\ell+2}\circ\dots\circ\mathcal{U}_{N+k}, (19)

where 𝒰N+ki(ρ):=(IiUN+ki)ρ(IiUN+ki)\mathcal{U}_{N+k-i}(\rho):=(I_{i}\otimes U_{N+k-i})\rho(I_{i}\otimes U_{N+k-i})^{\dagger} and the identity IiI_{i} acts on the radiated ii qubits. The quantum channel from the BH SS to the remaining BH SinS_{\rm in} and that to the new radiation SradS_{\rm rad}, denoted by SSin\mathcal{L}^{S\rightarrow S_{\rm in}} and SSrad\mathcal{L}^{S\rightarrow S_{\rm rad}}, are given by

SSin:=TrSradSSinSrad,\displaystyle\mathcal{L}^{S\rightarrow S_{\rm in}}:=\operatorname{Tr}_{S_{\rm rad}}\circ\mathcal{L}^{S\rightarrow S_{\rm in}S_{\rm rad}}, (20)
SSrad:=TrSinSSinSrad,\displaystyle\mathcal{L}^{S\rightarrow S_{\rm rad}}:=\operatorname{Tr}_{S_{\rm in}}\circ\mathcal{L}^{S\rightarrow S_{\rm in}S_{\rm rad}}, (21)

respectively. Note that they are complementary to each other since SSinSrad\mathcal{L}^{S\rightarrow S_{\rm in}S_{\rm rad}} is a unitary dynamics.

We then introduce another person Bob. He collects all the \ell qubits of the new radiation SradS_{\rm rad} and tries to recover the kk-qubit information originally stored in AA. He may additionally use the past radiation BradB_{\rm rad}. We assume that Bob knows the initial state |ξBinBrad|\xi\rangle^{B_{\rm in}B_{\rm rad}} and the whole dynamics SSinSrad\mathcal{L}^{S\rightarrow S_{\rm in}S_{\rm rad}}. In the recovery process, Bob tries to reproduce the state ΦAR\Phi^{AR} maximally entangled with the reference RR by applying a quantum channel 𝒟{\mathcal{D}} to the quantum systems he has, i.e., to the new and past radiations, SradS_{\rm rad} and BradB_{\rm rad}. If he succeeds, it implies that he can access the information source AA in the sense that, if the initial state of AA had been |ψ|\psi\rangle, Bob would be able to recover |ψ|\psi\rangle. Due to the no-cloning theorem [41], this implies that the information stored in AA is not left in the interior of the BH. Thus, the information in AA has been carried away by the radiation SradS_{\rm rad} of \ell qubits.

To quantitatively address whether Bob succeeds in recovering information of AA, we define and analyze a recovery error Δ\Delta. Let Φ^AR\hat{\Phi}^{AR} be the final state of the whole process, i.e.,

Φ^𝒟AR:=𝒟SradBradAABinSrad(ΦARξBinBrad).\hat{\Phi}^{AR}_{{\mathcal{D}}}:={\mathcal{D}}^{S_{\rm rad}B_{\rm rad}\rightarrow A}\circ\mathcal{L}^{AB_{\rm in}\rightarrow S_{\rm rad}}(\Phi^{AR}\otimes\xi^{B_{\rm in}B_{\rm rad}}). (22)

Since the goal of Bob is to establish the state maximally entangled with the reference RR, we define the recovery error by

Δ(ξ,):=min𝒟[1F(ΦAR,Φ^𝒟AR)],\Delta(\xi,\mathcal{L}):=\min_{{\mathcal{D}}}\bigl{[}1-F\bigl{(}\Phi^{AR},\hat{\Phi}^{AR}_{{\mathcal{D}}}\bigr{)}\bigr{]}, (23)

where the minimum is taken over all possible quantum channels 𝒟SradBradA{\mathcal{D}}^{S_{\rm rad}B_{\rm rad}\rightarrow A} applied by Bob onto the new and the past radiation, SradS_{\rm rad} and BradB_{\rm rad}. Note that 0Δ(ξ,)10\leq\Delta(\xi,\mathcal{L})\leq 1, and that the recovery error generally depends on the initial state ξ\xi of the initial BH BinB_{\rm in} and its dynamics \mathcal{L}.

4.2 Review of the HP protocol without symmetry

When the BH has no symmetry, the dynamics SSinSrad\mathcal{L}^{S\rightarrow S_{\rm in}S_{\rm rad}} is considered to be fully scrambling. This is formulated by choosing each internal unitary dynamics UN+kiU_{N+k-i} in 𝒰N+ki\mathcal{U}_{N+k-i} of the BH (see Eq. (19)) at Haar random. When this is the case, we denote the dynamics by Haar\mathcal{L}_{\rm Haar}. The recovery error Δ\Delta in this case is characterized by the min-entropy Hmin(Bin)ξH_{\rm min}(B_{\rm in})_{\xi} of the initial state of the BH BinB_{\rm in}: for the scrambling dynamics Haar\mathcal{L}_{\rm Haar} of the BH, the recovery error Δ\Delta satisfies [6, 42]

log2[Δ(ξ,Haar)]min{0,k+NHmin(Bin)ξ2},\log_{2}[\Delta(\xi,\mathcal{L}_{\rm Haar})]\\ \leq\min\bigl{\{}0,k+\frac{N-H_{\rm min}(B_{\rm in})_{\xi}}{2}-\ell\bigr{\}}, (24)

with high probability. This clearly shows that, when k+NHmin(Bin)ξ2\ell\geq k+\frac{N-H_{\rm min}(B_{\rm in})_{\xi}}{2}, the recovery error Δ(ξ,Haar)\Delta(\xi,\mathcal{L}_{\rm Haar}) decreases exponentially quickly. Note that the min-entropy Hmin(Bin)ξH_{\rm min}(B_{\rm in})_{\xi} is large if and only if the initial BH BinB_{\rm in} and the past radiation BradB_{\rm rad} are strongly entangled. Hence, in this framework, it is the entanglement between the initial BH BinB_{\rm in} and the past radiation BradB_{\rm rad} that determines the time scale for the information recovery to be possible.

For the pure BH, Hmin(Bin)ξ=0H_{\rm min}(B_{\rm in})_{\xi}=0, so that Δ(ξ,Haar)2N/2+k\Delta(\xi,\mathcal{L}_{\rm Haar})\leq 2^{N/2+k-\ell}. In contrast, for the mixed BH after the Page time [43], Hmin(Bin)ξ=NH_{\rm min}(B_{\rm in})_{\xi}=N, resulting in Δtot(ξ:Haar)2k\Delta_{tot}(\xi:\mathcal{L}_{\rm Haar})\leq 2^{k-\ell}. The latter case is particularly interesting since the recovery error does not depend on NN. Thus, no matter how large the initial BH is, Bob can recover the kk-qubit information when a little more than kk qubits are radiated.

The mechanism behind Eq. (24) is decoupling [44, 45, 46, 42] in the sense that HaarSSin(ΦARξBin)πSinπR\mathcal{L}_{\rm Haar}^{S\rightarrow S_{\rm in}}(\Phi^{AR}\otimes\xi^{B_{\rm in}})\approx\pi^{S_{\rm in}}\otimes\pi^{R}, which is induced by the fully scrambling dynamics of the BH. Decoupling is a basic concept in quantum information theory, which guarantees that information is encoded into good codewords, enabling Bob to retrieve information from the new radiation SradS_{\rm rad} easily.

4.3 The HP protocol with symmetry

Here, we describe how situation should be changed in the case of the Kerr BH with the ZZ-axial symmetry.

4.3.1 Symmetry, Hilbert spaces, and Quantum Information

When a quantum system has symmetry, the associated Hilbert space is accordingly decomposed. Since the axial symmetry is abelian, the Hilbert spaces A{\mathcal{H}}^{A}, Bin{\mathcal{H}}^{B_{\rm in}}, and S{\mathcal{H}}^{S} are decomposed into the subspaces invariant under the ZZ-axial symmetry:

A=κ=0kκA,Bin=μ=0NμBinandS=m=0N+kmS,{\mathcal{H}}^{A}=\bigoplus_{\kappa=0}^{k}\mathcal{H}_{\kappa}^{A},\ {\mathcal{H}}^{B_{\rm in}}=\bigoplus_{\mu=0}^{N}\mathcal{H}_{\mu}^{B_{\rm in}}\ \text{and}\ {\mathcal{H}}^{S}=\bigoplus_{m=0}^{N+k}\mathcal{H}_{m}^{S}, (25)

respectively. Note that the labels of the subspaces are based on the number of up-spins, but it can be readily transformed to the ZZ-axis AM. We denote the projection onto each subspace by ΠκS\Pi_{\kappa}^{S}, ΠμBin\Pi_{\mu}^{B_{\rm in}}, and ΠmS\Pi_{m}^{S}.

The symmetry also affects the quantum information stored in AA. As mentioned, quantum information in AA is represented by the maximally entangled state ΦAR\Phi^{AR} with the reference RR. The Hilbert space R{\mathcal{H}}^{R} is also decomposed according to the decomposition of A{\mathcal{H}}^{A}, namely, R=κκR{\mathcal{H}}^{R}=\bigoplus_{\kappa}\mathcal{H}^{R}_{\kappa}, where κR\mathcal{H}^{R}_{\kappa} is the range of TrA[ΠκAΦARΠκA]\operatorname{Tr}_{A}[\Pi_{\kappa}^{A}\Phi^{AR}\Pi_{\kappa}^{A}]. Using the projection ΠκR\Pi^{R}_{\kappa} onto κR\mathcal{H}^{R}_{\kappa}, we define

ΦdiagAR=κpκΦκAR,\Phi^{AR}_{\rm diag}=\sum_{\kappa}p_{\kappa}\Phi_{\kappa}^{AR}, (26)

with pκp_{\kappa} being Tr[(IAΠκR)ΦAR]\operatorname{Tr}[(I^{A}\otimes\Pi^{R}_{\kappa})\Phi^{AR}] and ΦκAR\Phi_{\kappa}^{AR} being (IAΠκR)ΦAR(IAΠκR)/pκ(I^{A}\otimes\Pi^{R}_{\kappa})\Phi^{AR}(I^{A}\otimes\Pi^{R}_{\kappa})/p_{\kappa}. Since ΦdiagAR\Phi_{\rm diag}^{AR} is invariant under the axial rotation of AA, this state contains the information of AA that is symmetry-invariant, leading us to call it symmetry-invariant information.

Accordingly, we define two errors in recovering information when the BH has symmetry. One is Δinv\Delta_{inv} for the symmetry-invariant information in ΦdiagAR\Phi^{AR}_{\rm diag}, and the other is Δtot\Delta_{tot} for the total information in ΦAR\Phi^{AR}. They are, respectively, defined by

Δinv(ξ,)=min𝒟[1F(ΦdiagAR,Φ^diag,𝒟AR)],\displaystyle\Delta_{inv}(\xi,\mathcal{L})=\min_{{\mathcal{D}}}\bigl{[}1-F\bigl{(}\Phi^{AR}_{\rm diag},\hat{\Phi}^{AR}_{{\rm diag},{\mathcal{D}}}\bigr{)}\bigr{]}, (27)
Δtot(ξ,)=min𝒟[1F(ΦAR,Φ^𝒟AR)].\displaystyle\Delta_{tot}(\xi,\mathcal{L})=\min_{{\mathcal{D}}}\bigl{[}1-F\bigl{(}\Phi^{AR},\hat{\Phi}^{AR}_{{\mathcal{D}}}\bigr{)}\bigr{]}. (28)

The recovery error Δtot\Delta_{tot} for the total information is the same as Δ\Delta in Eq. (23), but we below denote it by Δtot\Delta_{tot} to clearly distinguish it from Δinv\Delta_{inv}. Using a diagram is useful to represent the states Φ^diag,𝒟AR\hat{\Phi}^{AR}_{{\rm diag},{\mathcal{D}}} and Φ^𝒟AR\hat{\Phi}^{AR}_{{\mathcal{D}}} in Eqs. (27) and (28):

Φ^diag,𝒟AR= [Uncaptioned image],\displaystyle\hat{\Phi}^{AR}_{{\rm diag},{\mathcal{D}}}=\ \ \ {\begin{array}[]{c}\vbox{\vskip 0.75pt\hbox{\hskip 0.5pt\includegraphics[scale={0.25}]{TensorNetworkKerrBH0diag}}}\end{array}}, (30)
Φ^𝒟AR= [Uncaptioned image],\displaystyle\hat{\Phi}^{AR}_{{\mathcal{D}}}=\ \ \ {\begin{array}[]{c}\vbox{\vskip 0.75pt\hbox{\hskip 0.5pt\includegraphics[scale={0.25}]{TensorNetworkKerrBH0}}}\end{array}}, (32)

respectively. Here, Eq. (32) represents the state given by Eq. (22). The state defined by Eq. (30) is defined similarly by replacing Φ\Phi in Eq. (22) with Φdiag\Phi_{\rm diag}.

The states Φ^diag,𝒟AR\hat{\Phi}^{AR}_{{\rm diag},{\mathcal{D}}} and Φ^𝒟AR\hat{\Phi}^{AR}_{{\mathcal{D}}} are related to each other by the pinching map 𝒞R{\cal C}^{R} on RR, defined by 𝒞R(ρ):=κΠκRρΠκR{\cal C}^{R}(\rho):=\sum_{\kappa}\Pi_{\kappa}^{R}\rho\Pi_{\kappa}^{R} such as Φ^diag,𝒟AR=𝒞R(Φ^𝒟AR)\hat{\Phi}^{AR}_{{\rm diag},{\mathcal{D}}}={\cal C}^{R}(\hat{\Phi}^{AR}_{{\mathcal{D}}}). This is also represented as

Φ^diag,𝒟AR= [Uncaptioned image].\displaystyle\hat{\Phi}^{AR}_{{\rm diag},{\mathcal{D}}}=\ \ \ {\begin{array}[]{c}\vbox{\vskip 0.75pt\hbox{\hskip 0.5pt\includegraphics[scale={0.25}]{TensorNetworkKerrBH0anddiag}}}\end{array}}. (34)

4.3.2 Symmetry-preserving scrambling

Since the Kerr BH has a ZZ-axial symmetry, its internal unitary dynamics, namely, each UN+kiU_{N+k-i} in Eq. (19), should be in the form of

UN+ki=m=0N+kiUN+ki(m),U_{N+k-i}=\bigoplus_{m=0}^{N+k-i}U_{N+k-i}^{(m)}, (35)

where UN+ki(m)U_{N+k-i}^{(m)} acts on the subspace of N+kiN+k-i spins spanned by the states with mm up-spins. These unitaries form a group. We assume that each UN+ki(m)U_{N+k-i}^{(m)} is Haar random on each subspace, namely, UN+ki(m)U_{N+k-i}^{(m)} scrambles all quantum states with exactly mm up-spins. However, unlike the Haar scrambling, UN+kiU_{N+k-i} does not change the value of conserved quantity. We call such UN+kiU_{N+k-i} symmetry-preserving scrambling.

We denote the corresponding dynamics of the Kerr BH by Kerr\mathcal{L}_{\rm Kerr}. Using the unitarily invariant property of the Haar measure, it is straightforward to observe that the statistics induced by the dynamics KerrSSinSrad\mathcal{L}_{\rm Kerr}^{S\rightarrow S_{\rm in}S_{\rm rad}} is the same as that induced by a single application of a symmetry-preserving scrambling UKerrSU^{S}_{\rm Kerr} on the Kerr BH SS. Thus, in the following analysis, we consider UKerrSU^{S}_{\rm Kerr} instead of KerrSSinSrad\mathcal{L}_{\rm Kerr}^{S\rightarrow S_{\rm in}S_{\rm rad}}.

The symmetry-preserving scrambling unitary UKerrSU_{\rm Kerr}^{S} is formally defined by introducing a product of Haar measures: let 𝖧m{\sf H}_{m} be the Haar measure on the unitary group acting on the subspace with mm up-spins. Then, we define the symmetry-preserving scrambling UKerrSU_{\rm Kerr}^{S} by

UKerrS𝖧Kerr:=𝖧0×𝖧1××𝖧N+k.U_{\rm Kerr}^{S}\sim{\sf H}_{\rm Kerr}:={\sf H}_{0}\times{\sf H}_{1}\times\dots\times{\sf H}_{N+k}. (36)

The fact that the symmetry-preserving scrambling cannot change the conserved quantity immediately implies the absence of full decoupling in the Kerr BH, which was also pointed out in Ref. [14]. Since decoupling is the key technique used to derive the recovery error, Eq. (24), it is unclear at all whether and how information retrieval can be done in the presence of symmetry. The main goal of this paper is to clarify the recovery errors Δinv\Delta_{inv} and Δtot\Delta_{tot} for the symmetry-invariant and total information when the dynamics is Kerr\mathcal{L}_{\rm Kerr}, or equivalently UKerrU_{\rm Kerr}.

5 Information recovery –partial decoupling approach–

To investigate the recovery errors Δinv\Delta_{inv} and Δtot\Delta_{tot}, we use the partial decoupling approach [47], which we briefly summarize in Subsec. 5.1. We also provide a slight generalization. We then apply partial decoupling to the Kerr BH in Subsec. 5.2. We derive formulas for the upper bounds on the recover errors in Subsec. 5.4. Based on the formulas, we numerically investigate the recovery errors in Subsec. 5.5.

5.1 Partial decoupling theorem

The partial decoupling theorem is proven by two of the authors in Ref. [47]. We here explain its simplest form. In this subsection, the notation and the systems’ labeling, such as SS, RR, and EE, do not correspond to those in the HP protocol given in Sec. 4.

The situation of partial decoupling is as follows. Let SS and RR be quantum systems. We assume that the Hilbert space S{\mathcal{H}}^{S} has a direct-sum structure, such that

S=j𝒥jS.\mathcal{H}^{S}=\bigoplus_{j\in\mathcal{J}}\mathcal{H}^{S}_{j}. (37)

The dimensions are denoted as dimS=DS\dim{\mathcal{H}}^{S}=D_{S} and dimjS=dj\dim{\mathcal{H}}^{S}_{j}=d_{j}. Let ϱSR\varrho^{SR} be an initial state on SRSR. A random unitary USU^{S} in the form of j𝒥UjS\bigoplus_{j\in\mathcal{J}}U^{S}_{j} is first applied to SS, and then a given CPTP map SE{\mathcal{M}}^{S\rightarrow E} is applied. This leads to the state

SE(USϱSRUS).{\mathcal{M}}^{S\rightarrow E}(U^{S}\varrho^{SR}U^{S\dagger}). (38)

Despite the fact that this state is dependent on USU^{S}, the partial decoupling theorem states that, when US𝖧×:=𝖧1××𝖧JU^{S}\sim{\sf H}_{\times}:={\sf H}_{1}\times\dots\times{\sf H}_{J}, where 𝖧j{\sf H}_{j} is the Haar measure on the unitary group 𝖴(dj){\sf U}(d_{j}) on jS{\mathcal{H}}_{j}^{S}, the state is typically close to a fixed state independent of USU^{S} as far as a certain entropic condition is satisfied. For a later purpose, we slightly extend the situation, where the initial state ϱSR\varrho^{SR} is subnormalized, i.e., ϱSR𝒮(SR)\varrho^{SR}\in{\mathcal{S}}_{\leq}({\mathcal{H}}^{SR}), and the map SE{\mathcal{M}}^{S\rightarrow E} is a trace-non-increasing CP map.

The fixed (subnormalized) state is explicitly given by ΓER𝒮(SRE)\Gamma^{ER}\in\mathcal{S}_{\leq}(\mathcal{H}^{S^{*}\!RE}), which is constructed as

ΓER=j𝒥DSdjζjjEϱjjR,\Gamma^{ER}=\sum_{j\in\mathcal{J}}\frac{D_{S}}{d_{j}}\zeta^{E}_{jj}\otimes\varrho^{R}_{jj}, (39)

where

ζjjE=TrS[ΠjSζSE],\displaystyle\zeta^{E}_{jj}=\operatorname{Tr}_{S^{\prime}}\bigl{[}\Pi^{S^{\prime}}_{j}\zeta^{S^{\prime}E}\bigr{]}, (40)
ϱjjR=TrS[ΠjSϱSR],\displaystyle\varrho^{R}_{jj}=\operatorname{Tr}_{S}\bigl{[}\Pi^{S}_{j}\varrho^{SR}\bigr{]}, (41)

with ζSE=𝔍(SE)\zeta^{S^{\prime}E}=\mathfrak{J}({\mathcal{M}}^{S\rightarrow E}) being the Choi-Jamiołkowski representation of SE{\mathcal{M}}^{S\rightarrow E} (see Eq. (14)). Note that the state ΓER\Gamma^{ER} has correlation between EE and RR through the classical value jj and has no quantum correlation. Hence, we call the state partially decoupled.

The aforementioned entropic condition is given in terms of the conditional min-entropy of an extension ΓSRE𝒮(SRE)\Gamma^{S^{*}\!RE}\in\mathcal{S}_{\leq}(\mathcal{H}^{S^{*}\!RE}) of ΓER\Gamma^{ER} by a system S=SS′′S^{*}=S^{\prime}S^{\prime\prime}, where SS^{\prime} and S′′S^{\prime\prime} are replicas of the system SS:

ΓSER:=j,j𝒥DSdjdjζjjSEϱjjS′′R,\Gamma^{S^{*}ER}:=\sum_{j,j^{\prime}\in\mathcal{J}}\frac{D_{S}}{\sqrt{d_{j}d_{j^{\prime}}}}\zeta^{S^{\prime}E}_{jj^{\prime}}\otimes\varrho^{S^{\prime\prime}R}_{jj^{\prime}}, (42)

where

ζjjSE=ΠjSζSEΠjS,\displaystyle\zeta^{S^{\prime}E}_{jj^{\prime}}=\Pi^{S^{\prime}}_{j}\zeta^{S^{\prime}E}\Pi^{S^{\prime}}_{j^{\prime}}, (43)
ϱjjS′′R=ΠjSS′′ϱSR(ΠjSS′′).\displaystyle\varrho^{S^{\prime\prime}R}_{jj^{\prime}}=\Pi^{S\rightarrow S^{\prime\prime}}_{j}\varrho^{SR}(\Pi^{S\rightarrow S^{\prime\prime}}_{j^{\prime}})^{\dagger}. (44)

Here, ΠjSS′′=VSS′′ΠjS\Pi_{j}^{S\rightarrow S^{\prime\prime}}=V^{S\rightarrow S^{\prime\prime}}\Pi_{j}^{S}, with VSS′′V^{S\rightarrow S^{\prime\prime}} is a unitary that maps SS to S′′S^{\prime\prime}.

Using these subnormalized states, the partial decoupling theorem is stated as follows.

Theorem 1.

For any δ>0\delta>0, it holds that

SE(USϱSRUS)ΓER1212Hmin(S|ER)Γ+δ\bigl{\|}{\mathcal{M}}^{S\rightarrow E}\bigl{(}U^{S}\varrho^{SR}U^{S\dagger}\bigr{)}-\Gamma^{ER}\bigr{\|}_{1}\leq 2^{-\frac{1}{2}H_{\rm min}(S^{*}|ER)_{\Gamma}}+\delta (45)

with probability at least 1exp[δ2dmin48ϱS]1-\exp\bigl{[}-\frac{\delta^{2}d_{\rm min}}{48\|\varrho^{S}\|_{\infty}}\bigr{]} in terms of the choice of the unitary US𝖧×U^{S}\sim{\sf H}_{\times}, where dmin=minj𝒥{dj}d_{\rm min}=\min_{j\in{\mathcal{J}}}\{d_{j}\}.

The sub-normalized states ΓER\Gamma^{ER} and ΓSRE\Gamma^{S^{*}\!RE} have relatively simple expressions if we use a diagram. Let 𝒞𝒥SSS{\cal C}^{S\rightarrow SS^{*}}_{{\mathcal{J}}} be a CPTP map corresponding to a “𝒥{\mathcal{J}}-correlator” given by

𝒞𝒥SSS(ρS)=CSSSρS(CSSS),{\cal C}^{S\rightarrow SS^{*}}_{{\mathcal{J}}}(\rho^{S})=C^{S\rightarrow SS^{*}}\rho^{S}(C^{S\rightarrow SS^{*}})^{\dagger}, (46)

where CSSSC^{S\rightarrow SS^{*}} is the isometry given by

CSSS=j𝒥ΠjSS′′|ΦjSS.C^{S\rightarrow SS^{*}}=\sum_{j\in{\mathcal{J}}}\Pi_{j}^{S\rightarrow S^{\prime\prime}}\otimes|\Phi_{j}\rangle^{SS^{\prime}}. (47)

Here, |ΦjSS|\Phi_{j}\rangle^{SS^{\prime}} is the state maximally entangled only in the subspace jSjS{\mathcal{H}}_{j}^{S}\otimes{\mathcal{H}}_{j}^{S^{\prime}}. In other words, the 𝒥{\mathcal{J}}-correlator generates the maximally entangled state in the subspace jS′′jS{\mathcal{H}}_{j}^{S^{\prime\prime}}\otimes{\mathcal{H}}_{j}^{S^{\prime}} depending on the value of j𝒥j\in{\mathcal{J}} in the system SS, and swaps the system SS with S′′S^{\prime\prime}. Using this CPTP map, the subnormalized states ΓER\Gamma^{ER} and ΓSER\Gamma^{S^{*}ER} are expressed as follows:

ΓER= [Uncaptioned image],\displaystyle\Gamma^{ER}=\ \ \ {\begin{array}[]{c}\vbox{\vskip 0.72002pt\hbox{\hskip 0.48001pt\includegraphics[scale={0.24}]{TensorNetwork3}}}\end{array}}\ , (49)
ΓSER= [Uncaptioned image].\displaystyle\Gamma^{S^{*}ER}=\ \ \ {\begin{array}[]{c}\vbox{\vskip 0.72002pt\hbox{\hskip 0.48001pt\includegraphics[scale={0.24}]{TensorNetwork0}}}\end{array}}\ . (51)

In the above diagram, the black horizontal double bars represent the trace over the corresponding system, namely, SS^{*} in this case.

Theorem 1 implies that, if Hmin(S|ER)ΓH_{\rm min}(S^{*}|ER)_{\Gamma} is sufficiently large, the state SE(USϱSRUS){\mathcal{M}}^{S\rightarrow E}\bigl{(}U^{S}\varrho^{SR}U^{S\dagger}\bigr{)} is close to ΓER=TrS[ΓSER]\Gamma^{ER}=\operatorname{Tr}_{S^{*}}[\Gamma^{S^{*}ER}] with high probability. In the diagram representation, this can be rewritten as

 [Uncaptioned image] [Uncaptioned image],\displaystyle{\begin{array}[]{c}\vbox{\vskip 0.81001pt\hbox{\hskip 0.54001pt\includegraphics[scale={0.27}]{TensorNetwork2}}}\end{array}}\ \ \ \approx\ \ \ {\begin{array}[]{c}\vbox{\vskip 0.72002pt\hbox{\hskip 0.48001pt\includegraphics[scale={0.24}]{TensorNetwork3}}}\end{array}}\ , (54)

up to the approximation 2Hmin(S|ER)Γ/22^{-H_{\rm min}(S^{*}|ER)_{\Gamma}/2} in the trace norm with high probability.

When dmind_{\rm min} is small, such as dmin=1d_{\rm min}=1, Theorem 1 fails to provide a strong concentration since the probability in the statement becomes tiny for small δ\delta. When this is the case, we can set a “threshold” dimension dthd_{\rm th}.

Corollary 2.

Consider the same setting as in Theorem 1. Let dthd_{\rm th} be a positive integer, and ΠS\Pi_{\geq}^{S} be given by ΠS=j𝒥ΠjS\Pi_{\geq}^{S}=\sum_{j\in{\mathcal{J}}_{\geq}}\Pi_{j}^{S}, where 𝒥={j𝒥:djdth}{\mathcal{J}}_{\geq}=\{j\in{\mathcal{J}}:d_{j}\geq d_{\rm th}\}. If Tr[ϱSRΠS]1δ\operatorname{Tr}[\varrho^{SR}\Pi_{\geq}^{S}]\geq 1-\delta, then it holds for any δ>0\delta>0 that

SE(USϱSRUS)ΓER1212Hmin(S|ER)Γ1δ+δ+f(δ),\bigl{\|}{\mathcal{M}}^{S\rightarrow E}\bigl{(}U^{S}\varrho^{SR}U^{S\dagger}\bigr{)}-\Gamma^{ER}\bigr{\|}_{1}\\ \leq\frac{2^{-\frac{1}{2}H_{\rm min}(S^{*}|ER)_{\Gamma}}}{\sqrt{1-\delta}}+\delta+f(\delta), (55)

with probability at least 1exp[δ2dth48C]1-\exp\bigl{[}-\frac{\delta^{2}d_{\rm th}}{48C}\bigr{]}, where C=min{1,ϱS1δ}C=\min\{1,\frac{\|\varrho^{S}\|_{\infty}}{1-\delta}\}, and f(δ)=2δ+δ+δ1δf(\delta)=2\sqrt{\delta}+\delta+\frac{\delta}{1-\delta}.

The proofs of Theorem 1 and Corollary 2 are given in Appendices A and B, respectively.

5.2 Partial decoupling in the Kerr BH

We now apply the partial decoupling to the Kerr BH by identifying the dynamics KerrSSinSrad\mathcal{L}_{\rm Kerr}^{S\rightarrow S_{\rm in}S_{\rm rad}} with the symmetry-preserving scrambling UKerrS𝖧KerrU^{S}_{\rm Kerr}\sim{\sf H}_{\rm Kerr}, which is in the form of

UKerrS=m=0N+kUmS.U^{S}_{\rm Kerr}=\bigoplus_{m=0}^{N+k}U_{m}^{S}. (56)

Since this is the same form as the random unitary used in the partial decoupling theorem, we can directly apply Theorem 1 as well as Corollary 2.

We set ϱ\varrho and {\mathcal{M}} in Theorem 1 to ΦARξBin\Phi^{AR}\otimes\xi^{B_{\rm in}} and TrSrad\operatorname{Tr}_{S_{\rm rad}}, respectively. The system EE in Theorem 1 corresponds to SinS_{\rm in} in this case.

Using the doubled replica S=SS′′S^{*}=S^{\prime}S^{\prime\prime} of the BH SS, we have

 [Uncaptioned image] [Uncaptioned image],\displaystyle{\begin{array}[]{c}\vbox{\vskip 0.86998pt\hbox{\hskip 0.57999pt\includegraphics[scale={0.29}]{TensorNetworkBH1}}}\end{array}}\ \ \approx\ \ {\begin{array}[]{c}\vbox{\vskip 0.72002pt\hbox{\hskip 0.48001pt\includegraphics[scale={0.24}]{TensorNetworkBH2-2}}}\end{array}}, (59)

with high probability, where 𝒞AM{\cal C}_{\rm AM} is the AM-correlator that generates the state maximally entangled in a subspace. More explicitly, 𝒞AM{\cal C}_{\rm AM} is the conjugation map by the following isometry

CAMSSS=m=0N+kΠmSS′′|ΦmSS,C_{\rm AM}^{S\rightarrow SS^{*}}=\sum_{m=0}^{N+k}\Pi_{m}^{S\rightarrow S^{\prime\prime}}\otimes|\Phi_{m}\rangle^{SS^{\prime}}, (60)

where |ΦmSS|\Phi_{m}\rangle^{SS^{\prime}} is the state maximally entangled only in the subspace mSmS{\mathcal{H}}^{S}_{m}\otimes{\mathcal{H}}^{S^{\prime}}_{m}. According to Theorem 1, Eq. (59) holds up to 2Hmin(S|SinR)Γ/2\approx 2^{-H_{\rm min}(S^{*}|S_{\rm in}R)_{\Gamma}/2} in the trace norm, where

ΓSSinR\displaystyle\Gamma^{S^{*}S_{\rm in}R} = [Uncaptioned image].\displaystyle={\begin{array}[]{c}\vbox{\vskip 0.75pt\hbox{\hskip 0.5pt\includegraphics[scale={0.25}]{TensorNetworkBH2full}}}\end{array}}\ . (62)

Hence, if Hmin(S|SinR)Γ1H_{\rm min}(S^{*}|S_{\rm in}R)_{\Gamma}\gg 1, it is highly likely that the L.H.S. of Eq. (59) is partially decoupled, namely, the Kerr BH is partially decoupled from the reference RR.

5.3 Empirical smoothing of the conditional min-entropy

From Theorem 1, the probability for Eq. (59) to hold up to the approximation 212Hmin(S|ER)Γ+δ2^{-\frac{1}{2}H_{\rm min}(S^{*}|ER)_{\Gamma}}+\delta is at least 1exp[δ22kdmin/48]1-\exp[-\delta^{2}2^{k}d_{\rm min}/48], where we have used πAξBin2k\|\pi^{A}\otimes\xi^{B_{\rm in}}\|_{\infty}\leq 2^{-k}. Since dmin=1d_{\rm min}=1 for the axial symmetry, Theorem 1 fails to assure a high probability if δ<2k/2\delta<2^{-k/2}. This problem can be circumvented by using Corollary 2 instead of Theorem 1 and by setting a threshold dimension. However, the degree of approximation 212Hmin(S|ER)Γ2^{-\frac{1}{2}H_{\rm min}(S^{*}|ER)_{\Gamma}} turns out not to be tight in general, which typically happens when the conditional min-entropy is used. A smoothing is the technique that is exploited to obtain a better bound [48]. We here explain how the bound can be improved by the smoothing technique.

The most general way of smoothing is to use the smooth conditional min-entropy instead of the conditional min-entropy [49]. We here exploit a limited smoothing by ignoring “less probable” subspaces nSrad{\mathcal{H}}_{n}^{S_{\rm rad}} of the radiation SradS_{\rm rad}. This is because the smooth conditional min-entropy is computationally intractable.

To make the idea more precise, we consider the subnormalized state in which the ZZ-axis AM in the new radiation SradS_{\rm rad} is nn. We can explicitly compute the average weight pnp_{n}, namely, the trace of the subnormalized state, by taking the average over UKerrSU_{\rm Kerr}^{S}:

pn=12km=0N+kκ=0kχmκ(n)(N+kmn)(kκ)(N+km),p_{n}=\frac{1}{2^{k}}\sum_{m=0}^{N+k}\sum_{\kappa=0}^{k}\chi_{m-\kappa}\frac{\binom{\ell}{n}\binom{N+k-\ell}{m-n}\binom{k}{\kappa}}{\binom{N+k}{m}}, (63)

where χμ=Tr[ξBinΠμBin]\chi_{\mu}=\operatorname{Tr}[\xi^{B_{\rm in}}\Pi^{B_{\rm in}}_{\mu}] (see Appendix C). Using pnp_{n}, we define a probable set IϵI_{\epsilon} by

Iϵ={n[0,]:pnϵ},I_{\epsilon}=\{n\in[0,\ell]:p_{n}\geq\epsilon\}, (64)

for ϵ0\epsilon\geq 0. We call the subspaces nSrad{\mathcal{H}}_{n}^{S_{\rm rad}} for nIϵn\notin I_{\epsilon} rare events. Here, the parameter ϵ\epsilon defines the range of the probable set and rare events. We approximate, by using Theorem 1, only the subnormalized state in the subspace corresponding to the probable events, and count the part corresponding to the rare events as an additional error.

This procedure can be formulated by introducing Γ~(ϵ)\tilde{\Gamma}(\epsilon) corresponding to the probable events:

Γ~SSinR(ϵ):=ΠϵSradΓSSinRΠϵSrad,\displaystyle\tilde{\Gamma}^{S^{*}S_{\rm in}R}(\epsilon):=\Pi^{S_{\rm rad}}_{\epsilon}\Gamma^{S^{*}S_{\rm in}R}\Pi^{S_{\rm rad}}_{\epsilon}, (65)

where ΠϵSrad:=nIϵΠnSrad\Pi^{S_{\rm rad}}_{\epsilon}:=\sum_{n\in I_{\epsilon}}\Pi_{n}^{S_{\rm rad}}. The error from the rare events is then given by

w(ϵ)=1Tr[Γ~SSinR(ϵ)]=nIϵpn.w(\epsilon)=1-\operatorname{Tr}[\tilde{\Gamma}^{S^{*}S_{\rm in}R}(\epsilon)]=\sum_{n\notin I_{\epsilon}}p_{n}. (66)

Using these, we have the following Proposition (see Appendix D for the proof).

Proposition 3.

Let ϵ0\epsilon\geq 0. In the above setting, the probability that

 [Uncaptioned image] [Uncaptioned image],\displaystyle{\begin{array}[]{c}\vbox{\vskip 0.86998pt\hbox{\hskip 0.57999pt\includegraphics[scale={0.29}]{TensorNetworkBH1}}}\end{array}}\ \ \approx\ \ {\begin{array}[]{c}\vbox{\vskip 0.72002pt\hbox{\hskip 0.48001pt\includegraphics[scale={0.24}]{TensorNetworkBH2-2}}}\end{array}}, (69)

holds up to approximation Θξδ,ϵ\Theta_{\xi}^{\delta,\epsilon} in the trace norm is at least 1exp[δ2dmin(ϵ)48]1-\exp[-\frac{\delta^{2}d_{\rm min}(\epsilon)}{48}], where

Θξδ,ϵ(N,k,)=2112Hmin(S|SinR)Γ~(ϵ)+2w(ϵ)+2δ,\Theta^{\delta,\epsilon}_{\xi}(N,k,\ell)=2^{1-\frac{1}{2}H_{\rm min}(S^{*}|S_{\rm in}R)_{\tilde{\Gamma}(\epsilon)}}+2w(\epsilon)+2\delta, (70)

and dmin(ϵ)=min{(N+km+n):m[0,N+k],nIϵ}d_{\rm min}(\epsilon)=\min\bigl{\{}\binom{N+k}{m+n}:m\in[0,N+k-\ell],n\in I_{\epsilon}\bigr{\}}.

In the definition of Θξδ,ϵ\Theta^{\delta,\epsilon}_{\xi}, there is a term 2112Hmin(S|SinR)Γ~(ϵ)2^{1-\frac{1}{2}H_{\rm min}(S^{*}|S_{\rm in}R)_{\tilde{\Gamma}(\epsilon)}}. This term does not recover the non-smoothing bound 212Hmin(S|SinR)Γ2^{-\frac{1}{2}H_{\rm min}(S^{*}|S_{\rm in}R)_{\Gamma}} in the limit of ϵ0\epsilon\rightarrow 0. This may imply that the term could be further improved by factor 22.

5.4 Upper bounds on recovery errors

Based on the partial decoupling of the Kerr BH, i.e., Proposition 3 and using the standard technique in quantum information theory [44, 45, 46, 42], we now provide upper bounds on the recovery errors Δinv(ξ,Kerr)\Delta_{inv}(\xi,\mathcal{L}_{\rm Kerr}) and Δtot(ξ,Kerr)\Delta_{tot}(\xi,\mathcal{L}_{\rm Kerr}).

The key technique is the Uhlmann’s theorem, which states that, if any state ρAB\rho^{AB} and a pure state |σAC|\sigma\rangle^{AC} (both normalized) satisfy ρAσA1ε\|\rho^{A}-\sigma^{A}\|_{1}\leq\varepsilon, there exists a CPTP map 𝒟CB{\mathcal{D}}^{C\rightarrow B} such that

1εF(𝒟CB(|σσ|AC),ρAB).1-\varepsilon\leq F\bigl{(}{\mathcal{D}}^{C\rightarrow B}(|\sigma\rangle\langle\sigma|^{AC}),\rho^{AB}\bigr{)}. (71)

We apply this to Eq. (69). Since ξBin=TrBrad[|ξξ|BinBrad]\xi^{B_{\rm in}}=\operatorname{Tr}_{B_{\rm rad}}[|\xi\rangle\langle\xi|^{B_{\rm in}B_{\rm rad}}], we have

 [Uncaptioned image] [Uncaptioned image],\displaystyle{\begin{array}[]{c}\vbox{\vskip 0.81001pt\hbox{\hskip 0.54001pt\includegraphics[scale={0.27}]{TensorNetworkBH1-1}}}\end{array}}\approx\ \ {\begin{array}[]{c}\vbox{\vskip 0.72002pt\hbox{\hskip 0.48001pt\includegraphics[scale={0.24}]{TensorNetworkBH2full}}}\end{array}}\ , (74)

with approximation at least 1Θξδ,ϵ1-\Theta_{\xi}^{\delta,\epsilon} in terms of the fidelity, where 𝒟{\mathcal{D}} is a CPTP map from the radiations SradBradS_{\rm rad}B_{\rm rad} to SSS′′=SA′′Bin′′S^{*}\cong S^{\prime}S^{\prime\prime}=S^{\prime}A^{\prime\prime}B^{\prime\prime}_{\rm in}. Taking the partial trace over SBin′′S^{\prime}B^{\prime\prime}_{\rm in} as well as the remaining BH SinS_{\rm in}, we obtain

 [Uncaptioned image]\displaystyle{\begin{array}[]{c}\vbox{\vskip 0.81001pt\hbox{\hskip 0.54001pt\includegraphics[scale={0.27}]{TensorNetworkBH1-2}}}\end{array}}  [Uncaptioned image],\displaystyle\approx\ \ {\begin{array}[]{c}\vbox{\vskip 0.72002pt\hbox{\hskip 0.48001pt\includegraphics[scale={0.24}]{TensorNetworkBH4}}}\end{array}}, (77)
= [Uncaptioned image],\displaystyle=\ \ {\begin{array}[]{c}\vbox{\vskip 0.81001pt\hbox{\hskip 0.54001pt\includegraphics[scale={0.27}]{TensorNetworkBH5}}}\end{array}}, (79)

where 𝒞{\cal C} is the pinching map with respect to the ZZ-axis AM defined by 𝒞S(ρS)=mΠmSρSΠmS{\cal C}^{S}(\rho^{S})=\sum_{m}\Pi_{m}^{S}\rho^{S}\Pi_{m}^{S}. Note that S=ABinS=AB_{\rm in}. The last line simply follows from the fact that the AM-correlator 𝒞AM{\cal C}_{\rm AM} generates a normalized state depending on the value of the ZZ-axis AM in SS (see Eq. (60)), but it is traced out in Eq. (77).

To obtain the recovery error Δinv\Delta_{inv} for the symmetry-invariant information, we need to compute the fidelity to Φdiag=𝒞R(ΦAR)=κ=0kΠκRΦARΠκR\Phi_{\rm diag}={\cal C}^{R}(\Phi^{AR})=\sum_{\kappa=0}^{k}\Pi_{\kappa}^{R}\Phi^{AR}\Pi_{\kappa}^{R}. Using Eq. (79), we have

 [Uncaptioned image]= [Uncaptioned image], [Uncaptioned image]=ΦdiagAR.{\begin{array}[]{c}\vbox{\vskip 0.75pt\hbox{\hskip 0.5pt\includegraphics[scale={0.25}]{TensorNetworkBH1-2d}}}\end{array}}=\ {\begin{array}[]{c}\vbox{\vskip 0.75pt\hbox{\hskip 0.5pt\includegraphics[scale={0.25}]{TensorNetworkBH1-2dc}}}\end{array}},\\ \approx\ \ {\begin{array}[]{c}\vbox{\vskip 0.81001pt\hbox{\hskip 0.54001pt\includegraphics[scale={0.27}]{TensorNetworkBH5-2}}}\end{array}}=\Phi_{\rm diag}^{AR}. (80)

Recalling that the approximation is at least 1Θξδ,ϵ1-\Theta_{\xi}^{\delta,\epsilon} in the fidelity, we arrive at

Δinv(ξ,Kerr)1F(ΦdiagAR,Φ^diag,𝒟AR)Θξδ,ϵ.\Delta_{inv}(\xi,\mathcal{L}_{\rm Kerr})\leq 1-F\bigl{(}\Phi^{AR}_{\rm diag},\hat{\Phi}^{AR}_{{\rm diag},{\mathcal{D}}}\bigr{)}\leq\Theta_{\xi}^{\delta,\epsilon}. (81)

For the recovery error Δtot\Delta_{tot} for the total information, we can similarly show that

Δtot(ξ,Kerr)\displaystyle\Delta_{tot}(\xi,\mathcal{L}_{\rm Kerr}) KerrSSin(ΨSR)ΓSinπR1,\displaystyle\leq\bigl{\|}\mathcal{L}_{\rm Kerr}^{S\rightarrow S_{\rm in}}(\Psi^{SR})-\Gamma^{S_{\rm in}}\otimes\pi^{R}\bigr{\|}_{1}, (82)
KerrSSin(ΨSR)ΓSinR1\displaystyle\leq\bigl{\|}\mathcal{L}_{\rm Kerr}^{S\rightarrow S_{\rm in}}(\Psi^{SR})-\Gamma^{S_{\rm in}R}\bigr{\|}_{1}
+ΓSinRΓSinπR1,\displaystyle\hskip 42.67912pt+\bigl{\|}\Gamma^{S_{\rm in}R}-\Gamma^{S_{\rm in}}\otimes\pi^{R}\bigr{\|}_{1}, (83)
Θξδ,ϵ+ηξ,\displaystyle\leq\Theta_{\xi}^{\delta,\epsilon}+\eta_{\xi}, (84)

where ηξ\eta_{\xi} is given by

ηξ\displaystyle\eta_{\xi} =ΓSinRΓSinπR1\displaystyle=\|\Gamma^{S_{\rm in}R}-\Gamma^{S_{\rm in}}\otimes\pi^{R}\|_{1} (85)
=12kν=0N+kκ=0kFκ,ν(ξ)(N+kν)(kκ).\displaystyle=\frac{1}{2^{k}}\sum_{\nu=0}^{N+k-\ell}\sum_{\kappa=0}^{k}F_{\kappa,\nu}(\xi)\binom{N+k-\ell}{\nu}\binom{k}{\kappa}. (86)

Here,

Fκ,ν(ξ)=|m=0N+k(mν)(N+km)(χmκ12kκ=0k(kκ)χmκ)|,F_{\kappa,\nu}(\xi)\\ =\biggl{|}\sum_{m=0}^{N+k}\frac{\binom{\ell}{m-\nu}}{\binom{N+k}{m}}\biggl{(}\chi_{m-\kappa}-\frac{1}{2^{k}}\sum_{\kappa^{\prime}=0}^{k}\binom{k}{\kappa^{\prime}}\chi_{m-\kappa^{\prime}}\biggr{)}\biggr{|}, (87)

and χμ=Tr[ξBinΠμBin]\chi_{\mu}=\operatorname{Tr}[\xi^{B_{\rm in}}\Pi^{B_{\rm in}}_{\mu}]. See Appendix E for the derivation.

In summary, we have derived upper bounds on the recovery errors as follows.

Proposition 4.

Let ϵ,δ0\epsilon,\delta\geq 0. The recovery errors Δinv\Delta_{inv} and Δtot\Delta_{tot} for symmetry-invariant and total information satisfy

Δinv(ξ,Kerr)Θξδ,ϵ,\displaystyle\Delta_{inv}(\xi,\mathcal{L}_{\rm Kerr})\leq\Theta_{\xi}^{\delta,\epsilon}, (88)
Δtot(ξ,Kerr)Θξδ,ϵ+ηξ\displaystyle\Delta_{tot}(\xi,\mathcal{L}_{\rm Kerr})\leq\Theta_{\xi}^{\delta,\epsilon}+\eta_{\xi} (89)

with probability at least 1exp[δ2dmin(ϵ)48]1-\exp[-\frac{\delta^{2}d_{\rm min}(\epsilon)}{48}], where

Θξδ,ϵ(N,k,)=2112Hmin(S|SinR)Γ~(ϵ)+2w(ϵ)+2δ,\Theta^{\delta,\epsilon}_{\xi}(N,k,\ell)=2^{1-\frac{1}{2}H_{\rm min}(S^{*}|S_{\rm in}R)_{\tilde{\Gamma}(\epsilon)}}+2w(\epsilon)+2\delta, (90)

and dmin(ϵ)=min{(N+km+n):m[0,N+k],nIϵ}d_{\rm min}(\epsilon)=\min\bigl{\{}\binom{N+k}{m+n}:m\in[0,N+k-\ell],n\in I_{\epsilon}\bigr{\}}.

Since Proposition 4 holds for any ϵ0\epsilon\geq 0, we define Θξδ(N,k,)\Theta^{\delta}_{\xi}(N,k,\ell) by

Θξδ(N,k,):=minϵ0Θξδ,ϵ(N,k,),\Theta^{\delta}_{\xi}(N,k,\ell):=\min_{\epsilon\geq 0}\Theta^{\delta,\epsilon}_{\xi}(N,k,\ell), (91)

and investigate it in the following analysis. By definition, Θξδ(N,k,)\Theta^{\delta}_{\xi}(N,k,\ell) is the best possible upper bound of the recovery errors that can be obtained by the smoothing procedure described in Subsec. 5.3. We have numerically checked that this generally provides better bounds than those directly obtained from Theorem 1, namely, those without smoothing the entropy.

5.5 Numerical evaluation

The parameter δ\delta in Proposition 4 controls the trade-off between the upper bounds on the recovery errors and the probability for the symmetry-preserving scrambling Kerr\mathcal{L}_{\rm Kerr} to result in that errors. In the following numerical analysis, we consider only the term independent of δ\delta, that is,

Θξ(N,k,)=minϵ0{2112Hmin(S|SinR)Γ~(ϵ)+2w(ϵ)}.\displaystyle\Theta_{\xi}(N,k,\ell)=\min_{\epsilon\geq 0}\bigl{\{}2^{1-\frac{1}{2}H_{\rm min}(S^{*}|S_{\rm in}R)_{\tilde{\Gamma}(\epsilon)}}+2w(\epsilon)\bigr{\}}. (92)

This suffices since, in the large-NN limit, we can typically choose δ\delta such that Θξ(N,k,)δ\Theta_{\xi}(N,k,\ell)\gg\delta and, at the same time, that the probability is arbitrarily close to one. To see this, suppose that the ZZ-axis AM in BinB_{\rm in} is a constant fraction of NN. In this case, the dimension dmin(ϵ)d_{\rm min}(\epsilon) scales as Nc0N^{c_{0}\ell} (c0c_{0} is some constant). Hence, we can choose δ\delta to be values with a slightly greater scaling than Nc0/2N^{-c_{0}\ell/2}, such as Nc1/2N^{-c_{1}\ell/2} with c1<c0c_{1}<c_{0}, and make the probability close to one in the large-NN limit. On the other hand, as will be turned out later, Θξ(N,k,)\Theta_{\xi}(N,k,\ell) typically scales as 2c22^{-c_{2}\ell} (c2c_{2} is some constant). Since Θξ(N,k,)2c2Nc1/2=δ\Theta_{\xi}(N,k,\ell)\approx 2^{-c_{2}\ell}\gg N^{-c_{1}\ell/2}=\delta for sufficiently large NN, δ\delta is negligible and the probability is arbitrarily close to one.

To numerically compute Θξ(N,k,)\Theta_{\xi}(N,k,\ell), we need to simplify Hmin(S|SinR)Γ~(ϵ)H_{\rm min}(S^{*}|S_{\rm in}R)_{\tilde{\Gamma}(\epsilon)} that appears in Θξ\Theta_{\xi}. This can be done by using the property of the conditional min-entropy for the special type of states.

Let {jA}\{{\mathcal{H}}^{A}_{j}\} be mutually orthogonal subspaces of A{\mathcal{H}}^{A}, and πjA\pi_{j}^{A} be the completely mixed state on jA{\mathcal{H}}^{A}_{j}. For any state ΛABC\Lambda^{ABC} in the form of j=0JpjπjAρjBC\sum_{j=0}^{J}p_{j}\pi_{j}^{A}\otimes\rho_{j}^{BC}, where ρjBC𝒮(BC)\rho_{j}^{BC}\in\mathcal{S}(\mathcal{H}^{BC}) and {pj}\{p_{j}\} is a probability distribution, it holds that

2Hmin(AB|C)Λj=0Jpjdj2Hmin(B|C)ρj,2^{-H_{\rm min}(AB|C)_{\Lambda}}\leq\sum_{j=0}^{J}\frac{p_{j}}{d_{j}}2^{-H_{\rm min}(B|C)_{\rho_{j}}}, (93)

where dj=dimjAd_{j}=\dim{\mathcal{H}}^{A}_{j}. This immediately follows from the definition of the conditional min-entropy. See Appendix F for the derivation.

We also use the fact that, for a pure state |ψAB|\psi\rangle^{AB}, the conditional min-entropy is given by

Hmin(A|B)ψ\displaystyle H_{\rm min}(A|B)_{\psi} =2log[Tr[ψA]]\displaystyle=-2\log\bigl{[}\operatorname{Tr}[\sqrt{\psi^{A}}]\bigr{]} (94)
=2log[Tr[ψB]].\displaystyle=-2\log\bigl{[}\operatorname{Tr}[\sqrt{\psi^{B}}]\bigr{]}. (95)

Below, we provide upper bounds on the recovery errors for the pure and mixed Kerr BH. In both cases, we characterize the state ξ\xi of the initial Kerr BH BinB_{\rm in} by the expectation value LL of the ZZ-axis AM and the standard deviation δL\delta L, defined by

L=Tr[LZξ]\displaystyle L=\operatorname{Tr}[L_{Z}\xi] (96)
δL=(Tr[(LZL)2ξ])1/2,\displaystyle\delta L=\bigl{(}\operatorname{Tr}[(L_{Z}-L)^{2}\xi]\bigr{)}^{1/2}, (97)

where LZ=2i=1NZiL_{Z}=\frac{\hbar}{2}\sum_{i=1}^{N}Z_{i} with \hbar being the Plank constant and ZiZ_{i} being the Pauli-ZZ operator acting on the iith qubits. In the remaining of this paper, we set =1\hbar=1 except the last part of Subsec. 6.1, where we discuss a physical origin of information leakage.

Refer to caption
Figure 2: Recovery Errors from the radiation emitted by the pure Kerr BH. Upper bounds on Δinv(ξ:Kerr)\Delta_{inv}(\xi:\mathcal{L}_{\rm Kerr}) (dashed lines) and those on Δtot(ξ:Kerr)\Delta_{tot}(\xi:\mathcal{L}_{\rm Kerr}) (filled markers) for the pure Kerr BH are plotted. We particularly consider the initial pure states |ξBin|\xi\rangle^{B_{\rm in}} with various ZZ-axis AM LL and fluctuation δL\delta L. The figures (i) - (iv) show the cases for |L|=0,N/8,N/4|L|=0,N/8,N/4, and 3N/83N/8, respectively, while δL\delta L is chosen to be 0.1N0.1\sqrt{N} (red), 0.5N0.5\sqrt{N} (blue), 0.9N0.9\sqrt{N} (green), and 0.3N0.3N (brown) for each LL. For comparison, Δinv\Delta_{inv} for (L,δL)=(0,0)(L,\delta L)=(0,0) is plotted by a yellow dash-dotted line in each figure. The size NN of the initial Kerr BH and the size kk of the quantum information source are set to 500500 and 55, respectively. We observe that the error Δinv\Delta_{inv} for symmetry-invariant information decays exponentially quickly after =|Sin|\ell=|S_{\rm in}| exceeds a certain number, while the error Δtot\Delta_{tot} for the whole information typically has a plateau. The dependence of these features on LL and δL\delta L, as well as their physical mechanism, are discussed in the main text. Note however that, in any case, Δinv,Δtot1\Delta_{inv},\Delta_{tot}\approx 1 unless >(N+k)/2\ell>(N+k)/2, which is a consequence of the no-cloning property of quantum information [41].

5.5.1 The pure Kerr BH

Let us first consider the pure Kerr BH BinB_{\rm in} with the initial state in the form of

|ξ(L,δL)Bin=μ=0Nχμ(L,δL)|φμBin,|\xi(L,\delta L)\rangle^{B_{\rm in}}=\sum_{\mu=0}^{N}\sqrt{\chi_{\mu}(L,\delta L)}|\varphi_{\mu}\rangle^{B_{\rm in}}, (98)

where μ\mu stands for the number of up-spins in BinB_{\rm in}. We assume that {χμ(L,δL)}μ\{\chi_{\mu}(L,\delta L)\}_{\mu} are given by

χμ(L,δL)exp[(μLN/2)2δL2],\chi_{\mu}(L,\delta L)\propto\exp\biggl{[}-\frac{(\mu-L-N/2)^{2}}{\delta L^{2}}\biggr{]}, (99)

with a proper normalization such that μ=0Nχμ(L,δL)=1\sum_{\mu=0}^{N}\chi_{\mu}(L,\delta L)=1, and |φμBin|\varphi_{\mu}\rangle^{B_{\rm in}} is an arbitrary pure state in μBin{\mathcal{H}}^{B_{\rm in}}_{\mu}. Due to the symmetry-preserving scrambling, the recovery errors do not depend on the choice of |φμBin|\varphi_{\mu}\rangle^{B_{\rm in}}. Note that the expectation value of LZL_{Z} and its standard deviation for the state |ξ(L,δL)Bin|\xi(L,\delta L)\rangle^{B_{\rm in}} deviates from LL and δL\delta L, respectively, due to the discretization, but the deviations are negligibly small.

In this case, the subnormalized state Γ~SSinR(ϵ)\tilde{\Gamma}^{S^{*}S_{\rm in}R}(\epsilon) can be explicitly written as

Γ~L,δLSSinR(ϵ)=nIϵpnπnSradΦ~nSinSinS′′R(L,δL),\tilde{\Gamma}^{S^{*}S_{\rm in}R}_{L,\delta L}(\epsilon)=\sum_{n\in I_{\epsilon}}p_{n}\pi_{n}^{S^{\prime}_{\rm rad}}\otimes\tilde{\Phi}^{S_{\rm in}S_{\rm in}^{\prime}S^{\prime\prime}R}_{n}(L,\delta L), (100)

where {pn}\{p_{n}\} is the probability distribution given by Eq. (63), and

|Φ~n(L,δL)SinSinS′′Rm=0N+k1(N+km)(ΠmnSin|ΦSinSin)(ΠmSS′′(|ΦAR|ξBin)),|\tilde{\Phi}_{n}(L,\delta L)\rangle^{S_{\rm in}S^{\prime}_{\rm in}S^{\prime\prime}R}\propto\sum_{m=0}^{N+k}\sqrt{\frac{1}{\binom{N+k}{m}}}\bigl{(}\Pi_{m-n}^{S_{\rm in}}|\Phi\rangle^{S_{\rm in}S_{\rm in}^{\prime}}\bigr{)}\otimes\bigl{(}\Pi_{m}^{S\rightarrow S^{\prime\prime}}(|\Phi\rangle^{AR}\otimes|\xi\rangle^{B_{\rm in}})\bigr{)}, (101)

with a proper normalization constant. Here, we have used the notation that Πa\Pi_{a} is the zero operator if a<0a<0.

Using Eqs. (93) and (95), we obtain

Hmin(S|ER)Γ~(ϵ)klog[γpure(N,k,|ξ)],H_{\rm min}(S^{*}|ER)_{\tilde{\Gamma}(\epsilon)}\geq k-\log[\gamma_{\rm pure}(N,k,\ell|\xi)], (102)

where

γpure(N,k,|ξ)=nIϵ(m=0N+kκ=0kχmκ(N+km)(N+kmn)(kκ))2,\gamma_{\rm pure}(N,k,\ell|\xi)\\ =\sum_{n\in I_{\epsilon}}\biggl{(}\sum_{m=0}^{N+k}\sum_{\kappa=0}^{k}\sqrt{\frac{\chi_{m-\kappa}}{\binom{N+k}{m}}}\binom{N+k-\ell}{m-n}\binom{k}{\kappa}\biggr{)}^{2}, (103)

with χmκ=Tr[ξBinΠmκBin]\chi_{m-\kappa}=\operatorname{Tr}[\xi^{B_{\rm in}}\Pi^{B_{\rm in}}_{m-\kappa}]. This bound does not depend on the sign of LL. Namely, |ξ(±L,δL)|\xi(\pm L,\delta L)\rangle result in the same bound. This is naturally expected since the rotation direction of the Kerr BH should not affect any features of the protocol.

From Eqs. (92) and (102), we obtain

Θξ(N,k,)minϵ0{21k/2γpure(N,k,|ξ)+2w(ϵ)}.\Theta_{\xi}(N,k,\ell)\leq\\ \min_{\epsilon\geq 0}\bigl{\{}2^{1-k/2}\sqrt{\gamma_{\rm pure}(N,k,\ell|\xi)}+2w(\epsilon)\bigr{\}}. (104)

Since w(ϵ)=nIϵpnw(\epsilon)=\sum_{n\notin I_{\epsilon}}p_{n}, we can numerically evaluate this bound, from which we obtain upper bounds on Δinv\Delta_{inv} and Δtot\Delta_{tot}.

In Fig. 2, upper bounds on Δinv\Delta_{inv} and Δtot\Delta_{tot}, i.e.,

Δinv(ξ,Kerr)Θξ,\displaystyle\Delta_{inv}(\xi,\mathcal{L}_{\rm Kerr})\leq\Theta_{\xi}, (105)
Δtot(ξ,Kerr)Θξ+ηξ\displaystyle\Delta_{tot}(\xi,\mathcal{L}_{\rm Kerr})\leq\Theta_{\xi}+\eta_{\xi} (106)

are provided as functions of the number \ell of qubits in the new radiation SradS_{\rm rad} for various LL and δL\delta L. We observe that Δinv\Delta_{inv} starts decaying exponentially after a certain number of \ell. The number, as well as the decaying speed, strongly depends on both |L||L| and δL\delta L. To characterize this, we introduce Δ(L,δL)\ell_{\Delta}(L,\delta L) for Δ>0\Delta>0 by

Δ(L,δL)=min{[0,N+k]|ΘξΔ}.\ell_{\Delta}(L,\delta L)=\min\bigl{\{}\ell\in[0,N+k]|\Theta_{\xi}\leq\Delta\bigr{\}}. (107)

Then, it is observed from Fig. 2 that Δ(L,δL)Δ(0,0)\ell_{\Delta}(L,\delta L)\geq\ell_{\Delta}(0,0) for any Δ0\Delta\geq 0. This implies that the Kerr BH has a delay in the onset of releasing information compared to the trivial case (L=δL=0L=\delta L=0), which is given by the yellow dotted lines in Fig. 2. The delay is especially large when LL is large (see figure (iv)) or when δL=Θ(N)\delta L=\Theta(N) (see brown plots), indicating that the symmetry-invariant information cannot be recovered from the radiation if either LL or δL\delta L is extremely large.

We also observe that Δtot\Delta_{tot} behaves differently depending on whether δL=O(N)\delta L=O(\sqrt{N}) or O(N)O(N). When δL=O(N)\delta L=O(\sqrt{N}), Δtot\Delta_{tot} first behaves similarly to Δinv\Delta_{inv}. However, soon after that, the decreasing of the error becomes slow, and Δtot\Delta_{tot} remains at a non-negligible value until N\ell\approx N. This indicates that a part of the information remains in the Kerr BH until the last moment. We call such residual information information remnant. From the plots for different LL but same δL\delta L, we also observe that the amount of information remnant is independent of LL. When δL=O(N)\delta L=O(N), ΔtotΔinv\Delta_{tot}\approx\Delta_{inv} for any \ell, implying that there is no information remnant.

5.5.2 The mixed Kerr BH

Refer to caption
Figure 3: Recovery Errors from the radiation emitted by the mixed Kerr BH. Upper bounds on Δinv(ξ:Kerr)\Delta_{inv}(\xi:\mathcal{L}_{\rm Kerr}) (dashed lines) and those on Δtot(ξ:Kerr)\Delta_{tot}(\xi:\mathcal{L}_{\rm Kerr}) (filled markers) for the mixed Kerr BH are plotted. The mixed Kerr BH BinB_{\rm in} is entangled with the past radiation BradB_{\rm rad}. We consider a family of specific entangled states with various ZZ-axis AM LL and fluctuation δL\delta L. The figures (i) - (iv) show the bounds for |L|=0,N/8,N/4|L|=0,N/8,N/4, and 3N/83N/8, respectively, where ±L\pm L leads to the same result as explained in the caption of Fig. 2, while δL\delta L is chosen to be 0.1N0.1\sqrt{N} (red), 0.5N0.5\sqrt{N} (blue), 0.9N0.9\sqrt{N} (green), and 0.3N0.3N (brown). The recovery error Δinv\Delta_{inv} for (L,δL)=(0,0)(L,\delta L)=(0,0) is also plotted by a yellow dash-dotted line for comparison. The size NN of the initial Kerr BH is set to 500500, and the size kk of the quantum information source to 55. Similarly to the pure Kerr BH, we observe that both Δinv(ξ:Kerr)\Delta_{inv}(\xi:\mathcal{L}_{\rm Kerr}) and Δtot(ξ:Kerr)\Delta_{tot}(\xi:\mathcal{L}_{\rm Kerr}) start decreasing at almost same timing, and that only Δtot\Delta_{tot} stops decreasing soon after, except the one for δL=0.3N\delta L=0.3N.

We next consider the mixed Kerr BH, which is initially entangled with the past radiation BradB_{\rm rad}. For simplicity, we only consider the following entangled states:

|ξ(L,δL)BinBrad=μ=0Nχμ(L,δL)|ΦμBinBrad,|\xi(L,\delta L)\rangle^{B_{\rm in}B_{\rm rad}}=\sum_{\mu=0}^{N}\sqrt{\chi_{\mu}(L,\delta L)}|\Phi_{\mu}\rangle^{B_{\rm in}B_{\rm rad}}, (108)

where {χμ}μ\{\chi_{\mu}\}_{\mu} are given by Eq. (99), and |ΦμBinBrad|\Phi_{\mu}\rangle^{B_{\rm in}B_{\rm rad}} is the maximally entangled state in μBinμBrad{\mathcal{H}}^{B_{\rm in}}_{\mu}\otimes{\mathcal{H}}^{B_{\rm rad}}_{\mu}.

In this case, the state ΓSSinR\Gamma^{S^{*}S_{\rm in}R} reduces to

ΓSSinR=n=0μ=0Npn,μπnSradπμBin′′Ψ~n,μSinSinA′′R,\Gamma^{S^{*}S_{\rm in}R}=\sum_{n=0}^{\ell}\sum_{\mu=0}^{N}p_{n,\mu}\pi_{n}^{S^{\prime}_{\rm rad}}\otimes\pi_{\mu}^{B^{\prime\prime}_{\rm in}}\otimes\tilde{\Psi}_{n,\mu}^{S_{\rm in}S^{\prime}_{\rm in}A^{\prime\prime}R}, (109)

where pn,μp_{n,\mu} and the normalized pure state Ψ~n,μSinSinA′′R\tilde{\Psi}_{n,\mu}^{S_{\rm in}S^{\prime}_{\rm in}A^{\prime\prime}R} are given by

pn,μ=χμ(L,δL)(n)2km=0N+k(N+kmn)(kmμ)(N+km),\displaystyle p_{n,\mu}=\frac{\chi_{\mu}(L,\delta L)\binom{\ell}{n}}{2^{k}}\sum_{m=0}^{N+k}\frac{\binom{N+k-\ell}{m-n}\binom{k}{m-\mu}}{\binom{N+k}{m}}, (110)

and

|Ψ~n,μSinSinA′′Rm=0N+k1(N+km)(ΠmnSin|ΦSinSin)(ΠmμAA′′|ΦAR),|\tilde{\Psi}_{n,\mu}\rangle^{S_{\rm in}S^{\prime}_{\rm in}A^{\prime\prime}R}\propto\\ \sum_{m=0}^{N+k}\frac{1}{\sqrt{\binom{N+k}{m}}}\bigl{(}\Pi^{S_{\rm in}}_{m-n}|\Phi\rangle^{S_{\rm in}S^{\prime}_{\rm in}}\bigr{)}\otimes\bigl{(}\Pi^{A\rightarrow A^{\prime\prime}}_{m-\mu}|\Phi\rangle^{AR}\bigr{)}, (111)

respectively. Note that μpn,μ=pn\sum_{\mu}p_{n,\mu}=p_{n}, where pnp_{n} is given by Eq. (63).

Using Eqs. (93) and (95), we obtain

Hmin(S|ER)Γ~(ϵ)klog[γmixed(N,k,|ξ)]H_{\rm min}(S^{*}|ER)_{\tilde{\Gamma}(\epsilon)}\geq k-\log\bigl{[}\gamma_{\rm mixed}(N,k,\ell|\xi)\bigr{]} (112)

where

γmixed(N,k,|ξ):=μ=0Nχμ(L,δL)(Nμ)nIϵ(m=0N+k1(N+km)(N+kmn)(kmμ))2.\gamma_{\rm mixed}(N,k,\ell|\xi):=\sum_{\mu=0}^{N}\frac{\chi_{\mu}(L,\delta L)}{\binom{N}{\mu}}\sum_{n\in I_{\epsilon}}\\ \biggl{(}\sum_{m=0}^{N+k}\sqrt{\frac{1}{\binom{N+k}{m}}}\binom{N+k-\ell}{m-n}\binom{k}{m-\mu}\biggr{)}^{2}. (113)

This leads to

Θξ(N,k,)minϵ0{21k/2γmixed(N,k,|ξ)+2w(ϵ)},\Theta_{\xi}(N,k,\ell)\\ \leq\min_{\epsilon\geq 0}\bigl{\{}2^{1-k/2}\sqrt{\gamma_{\rm mixed}(N,k,\ell|\xi)}+2w(\epsilon)\bigr{\}}, (114)

and, we obtain upper bounds on Δinv\Delta_{inv} and Δtot\Delta_{tot}. We again note that ±L\pm L provide the same bounds.

    δL=O(N)\delta L=O(\sqrt{N})     δL=O(N)\delta L=O(N)
Delay for the pure Kerr BH for any LL O(N)O(\sqrt{N}) O(N)O(N)
Delay for the mixed Kerr BH for small LL O(1)O(1) O(N)O(\sqrt{N})?
for large LL O(1)O(1) O(N)O(N)?
Table 2: A summary of the delay of information leakage for the pure and mixed Kerr BHs, which initially have the ZZ-axis AM LL and the fluctuation δL\delta L. The results are obtained from the numerical analysis based on partial decoupling, where we varies NN up to 500500. The order is only in terms of N=|Bin|N=|B_{\rm in}|. In the case of the mixed Kerr BH with small LL and δL=O(N)\delta L=O(N), it was hard to decide the scaling from our numerics.

The upper bounds on the recovery errors are numerically plotted in Fig. 3 as functions of =|Srad|\ell=|S_{\rm rad}|. Similarly to the pure Kerr BH, the plots are for various LL and δL\delta L. We observe that Δinv\Delta_{inv} decays in a manner similar to the pure Kerr BH. That is, the recovery error starts decreasing after a certain number of \ell, as well as the onset of decreasing has a delay compared to the trivial case (L=δL=0L=\delta L=0, yellow dotted lines in Fig. 3). While the delay is not so large for small LL and δL\delta L, it becomes substantially large when |L||L| is large or δL=Θ(N)\delta L=\Theta(N).

The recovery error Δtot\Delta_{tot} for the total information also behaves similarly to the pure Kerr BH, and it remains at a non-zero value until N\ell\approx N. This implies that, even for the mixed Kerr BH, the information remnant exists. It turns out that the amount of information remnant depends only on δL\delta L and is the same as that for the pure Kerr BH for the same δL\delta L, which will be elaborated on later.

The significant difference from the pure Kerr BH is that the timing at which the errors start decreasing can be much earlier. It indeed ranges widely from O(k)O(k) to O(N)O(N) depending on the ZZ-axis AM LL and on its fluctuation δL\delta L of the initial Kerr BH.

5.5.3 Scaling of the delay with respect to NN

We also provide numerical results about how the delay Δ(L,δL)\ell_{\Delta}(L,\delta L) scales with respect to the size NN of the initial Kerr BH BinB_{\rm in} for various ZZ-axis AMs LL and their fluctuations δL\delta L. The result is summarized in Tab. 2 for the pure and mixed Kerr BHs. See Appendix G for the details.

6 Information recovery –macroscopic physics approach–

In Sec. 5, we have numerically shown that the delay and the information remnant are the consequences of the symmetry of the Kerr BH. In this section, we analyze them from different perspectives. This lead to great insights into their physical origins.

We consider the delay in Subsec. 6.1 and the information remnant in Subsec. 6.2.

6.1 Delay of information leakage and themodynamics

To interpret the delay of information leakage, we propose a new concept that we call clipping of entanglement and show that it reproduces both the result for the BH without symmetry and the delay of information leakage for the Kerr BH. As we will see, the clipping argument provides a concise condition for the information leakage to occur only in terms of dimensions of the Hilbert spaces.

Before we start, we emphasize that the argument based on the entanglement clipping is applicable not only for the axial symmetry but for any symmetry with extensive conserved quantity.

6.1.1 Clipping of Entanglement

One of the significant features of the scrambling dynamics is that it generates nearly maximal entanglement between subsystems [43, 50, 51]. We argue that such a maximal entanglement is responsible also for the quick information leakage.

For clarity, we first consider the BH without symmetry. Moreover, instead of directly considering the quantum information source ΦAR\Phi^{AR}, i.e., the maximally entangled state between AA and the reference RR, we consider the situation where the reference RR is measured in an arbitrary basis to obtain a kk-bit outcome. It is known that if the kk-bit outcome can be inferred from the radiations BradB_{\rm rad} and SradS_{\rm rad}, then the maximally entangled state ΦAR\Phi^{AR} can be reproduced [52], that is, the quantum information in AA is recovered.

When the BH has no symmetry, its dynamics is fully scrambling and generates nearly maximal entanglement between SradS_{\rm rad} and SinS_{\rm in}. This implies that the marginal state in the smaller subsystem should be nearly completely mixed. In contrast, the support of the marginal state in the larger subsystem needs to be “clipped” at the dimension of the smaller one, because any bipartite pure state should have an equal size, known as the Schmidt rank, in the two subsystems [53].

The information recovery can be understood from the entanglement clipping in the larger subsystem. In the case of the pure BH, no information is obtained from the radiation when |Srad||Sin||S_{\rm rad}|\leq|S_{\rm in}| since no clipping occurs in the new radiation SradS_{\rm rad}. Note that all the eigenvalues of the marginal state in SradS_{\rm rad} are nearly the same, from which no information would be retrieved. When |Srad||Sin||S_{\rm rad}|\geq|S_{\rm in}|, the clipping takes place in SradS_{\rm rad}, namely, the marginal state in SradS_{\rm rad} is released from the full rank state. In the case of the mixed BH, the entanglement clipping in SradS_{\rm rad} sets in much earlier since the BH is initially already on the verge of clipping due to the entanglement with the past radiation BradB_{\rm rad}, which is under control of the decoding person Bob.

To capture the information-theoretic consequence of the entanglement clipping in the new radiation SradS_{\rm rad}, we introduce a degree of clipping CC in Srad{\mathcal{H}}^{S_{\rm rad}}:

C(Srad):=H(Bin)ξ+log[dimSraddimSin].C({\mathcal{H}}^{S_{\rm rad}}):=H(B_{\rm in})_{\xi}+\log\biggl{[}\frac{\dim{\mathcal{H}}^{S_{\rm rad}}}{\dim{\mathcal{H}}^{S_{\rm in}}}\biggr{]}. (115)

Note that the degree of clipping C(Srad)C({\mathcal{H}}^{S_{\rm rad}}) can be negative. For instance, C(Srad)=(N+k)C({\mathcal{H}}^{S_{\rm rad}})=-(N+k) when |Srad|=0|S_{\rm rad}|=0 and H(Bin)ξ=0H(B_{\rm in})_{\xi}=0. When C(Srad)>kC({\mathcal{H}}^{S_{\rm rad}})>k, the marginal state in the new radiation SradS_{\rm rad} is sufficiently clipped so that the 2k2^{k} states, corresponding to all the possible 2k2^{k} outcomes in RR, can be fit in the marginal state without significant overlaps each other. Thus, we propose that the condition for the information recovery is given by

C(Srad)>k.C({\mathcal{H}}^{S_{\rm rad}})>k. (116)

By explicitly writing down the dimensions, we obtain

>k+NH(Bin)ξ2,\ell>k+\frac{N-H(B_{\rm in})_{\xi}}{2}, (117)

as a condition for the information recovery to be possible. Note that this agrees well with the result of the HP protocol without symmetry, namely, Eq. (24).

We can apply the same argument to the Kerr BH. In this case, it is of crucial importance that the symmetry-preserving scrambling in the form of mUmS\bigoplus_{m}U^{S}_{m}, where UmSU^{S}_{m} is the Haar scrambling in the subspace with mm up-spins, generates entanglement between SinS_{\rm in} and SradS_{\rm rad} only in the subspaces. This introduces a special structure of entanglement between them. To be more precise, we consider the initial Kerr BH BinB_{\rm in} with a fixed ZZ-axis AM L=λNL=\lambda N and δL=0\delta L=0, where λ[1/2,1/2]\lambda\in[-1/2,1/2]. The number of up-spins in BinB_{\rm in} is λN+N/2\lambda N+N/2. Due to the symmetry-preserving scrambling, the subspace nSrad{\mathcal{H}}_{n}^{S_{\rm rad}} with nn up-spins in SradS_{\rm rad} is entangled only with nSin{\mathcal{H}}_{n^{\prime}}^{S_{\rm in}} satisfying n+nλN+(N+k)/2n+n^{\prime}\approx\lambda N+(N+k)/2.

Taking this constraint into account, the entanglement clipping should be considered in each subspace nSrad{\mathcal{H}}_{n}^{S_{\rm rad}} separately. We thus define the degree of clipping in nSrad{\mathcal{H}}_{n}^{S_{\rm rad}} by

C(nSrad):=H(Bin)ξ+log[dimnSraddimλN+(N+k)/2nSin],C({\mathcal{H}}_{n}^{S_{\rm rad}}):=H(B_{\rm in})_{\xi}+\log\biggl{[}\frac{\dim{\mathcal{H}}^{S_{\rm rad}}_{n}}{\dim{\mathcal{H}}^{S_{\rm in}}_{\lambda N+(N+k)/2-n}}\biggr{]}, (118)

and require that

C(nSrad)>k,C({\mathcal{H}}_{n}^{S_{\rm rad}})>k, (119)

for most nn.

We here emphasize that we have not used any property of the ZZ-axial symmetry of the Kerr BH except that it is additive. Thus, the above clipping condition should be valid for any BH with extensive conserved quantity.

If Eq. (119) is not satisfied for certain nn, the information in the corresponding subspace should be counted as recovery error. This point can be elaborated by considering the probability distribution W(n)W(n) over the number nn of spins in SradS_{\rm rad} induced by the random choice of SradS_{\rm rad} from SS. The probability is given by

W(n)=dimλN+(N+k)/2nSindimnSraddimλN+(N+k)/2S.W(n)=\frac{\dim{\mathcal{H}}_{\lambda N+(N+k)/2-n}^{S_{\rm in}}\dim{\mathcal{H}}_{n}^{S_{\rm rad}}}{\dim{\mathcal{H}}_{\lambda N+(N+k)/2}^{S}}. (120)

With respect to this probability distribution, we define probable values of nn and require for Eq. (119) to hold for any probable nn. We count the total probability over non-probable values of nn as recovery errors. This is similar to the empirical smoothing we exploited in Subsec. 5.3.

6.1.2 Comparison between entanglement clipping and partial decoupling

To check the validity of the argument of entanglement clipping, let us compare the condition (119) with the partial decoupling result. To this end, we need to specify the probable set of nn. We define the probable set by the set of nn around its average n\langle n\rangle under the probability distribution of Eq. (120). The average is explicitly given by

n=(LN+k+1/2).\langle n\rangle=\biggl{(}\frac{L}{N+k}+1/2\biggr{)}\ell. (121)

We require that Eq. (119) holds for all nn such that

|nn|cδn2,|n-\langle n\rangle|\leq c\sqrt{\langle\delta n^{2}\rangle}, (122)

where δn:=nn\delta n:=n-\langle n\rangle. Here, c>0c>0 is a parameter related to the recovery error since taking a larger cc means that we require Eq. (119) for a wider range of nn, resulting in less error. We then define the number ^c(L)\hat{\ell}_{c}(L) by the minimum of =|Srad|\ell=|S_{\rm rad}| for the clipping condition C(nSrad)>kC({\mathcal{H}}_{n}^{S_{\rm rad}})>k to hold for any nn satisfying Eq. (122).

The quantity ^c(L)\hat{\ell}_{c}(L) should correspond to the Δ(L,0)\ell_{\Delta}(L,0), defined based on the partial decoupling by Eq. (107). In the following, we show that ^c(L)Δ(L,0)\hat{\ell}_{c}(L)\approx\ell_{\Delta}(L,0), which indicates that the clipping argument results in nearly the same prediction as the partial decoupling.

To this end, we provide an approximate expression of ^c(L)\hat{\ell}_{c}(L). The dimensions in Eq. (118) are given by

dimnSrad=(n),\displaystyle\dim{\mathcal{H}}_{n}^{S_{\rm rad}}=\binom{\ell}{n}, (123)
dimλN+(N+k)/2nSin=(N+kλN+(N+k)/2n).\displaystyle\dim{\mathcal{H}}_{\lambda N+(N+k)/2-n}^{S_{\rm in}}=\binom{N+k-\ell}{\lambda N+(N+k)/2-n}. (124)

These are well-approximated by a function s(λ):=(1/2λ)log[1/2λ](1/2+λ)log[1/2+λ]s(\lambda):=-(1/2-\lambda)\log[1/2-\lambda]-(1/2+\lambda)\log[1/2+\lambda]. We then obtain

log[dimnSraddimλN+(N+k)/2nSin]=s(n12)(N+k)s(λN+(N+k)/2nN+k12).\log\biggl{[}\frac{\dim{\mathcal{H}}_{n}^{S_{\rm rad}}}{\dim{\mathcal{H}}_{\lambda N+(N+k)/2-n}^{S_{\rm in}}}\biggr{]}=\ell\ \!s\biggl{(}\frac{n}{\ell}-\frac{1}{2}\biggr{)}-(N+k-\ell)\ \!s\biggl{(}\frac{\lambda N+(N+k)/2-n}{N+k-\ell}-\frac{1}{2}\biggr{)}. (125)

We can further simplify the expression by assuming kNk\ll N and expanding nn around its average n(λ+1/2)\langle n\rangle\approx(\lambda+1/2)\ell as

n(λ+1/2)+δn.n\approx(\lambda+1/2)\ell+\delta n. (126)

It is straightforward to show

log[dimnSraddimλN+(N+k)/2nSin](N+k2)s(λ)+2δn|s(λ)|.\log\biggl{[}\frac{\dim{\mathcal{H}}_{n}^{S_{\rm rad}}}{\dim{\mathcal{H}}_{\lambda N+(N+k)/2-n}^{S_{\rm in}}}\biggr{]}\\ \approx-(N+k-2\ell)\ \!s(\lambda)+2\delta n|s^{\prime}(\lambda)|. (127)

We also introduce the initial degree of clipping CiniC_{\rm ini}, i.e., the degree of clipping when |Srad|=0|S_{\rm rad}|=0. It is approximately given by

CiniH(Bin)ξ(N+k)s(λ)<0.C_{\rm ini}\approx H(B_{\rm in})_{\xi}-(N+k)s(\lambda)<0. (128)

Using these, we arrive at

C(nSrad)Cini+2s(λ)+2δn|s(λ)|.C({\mathcal{H}}^{S_{\rm rad}}_{n})\approx C_{\rm ini}+2\ell s(\lambda)+2\delta n|s^{\prime}(\lambda)|. (129)

As mentioned, we require C(nSrad)>kC({\mathcal{H}}^{S_{\rm rad}}_{n})>k for all δn\delta n such that |δn|cδn2|\delta n|\leq c\sqrt{\langle\delta n^{2}\rangle}. The standard deviation δn2\sqrt{\langle\delta n^{2}\rangle} can be also rephrased in terms of s(λ)s(\lambda). From Eq. (120), we have

log[W(n)]\displaystyle\log[W(n)] =12(N+k)(N+k)|s′′(λ)|δn2+O(δn3).\displaystyle=-\frac{1}{2}\frac{(N+k)}{\ell(N+k-\ell)}|s^{\prime\prime}(\lambda)|\delta n^{2}+O(\delta n^{3}). (130)

Since W(n)W(n) can be approximated by a Gaussian distribution when 1kN1\ll k\ll N, the variance δn2\langle\delta n^{2}\rangle is approximately given by

δn2(1N+k)|s′′(λ)|.\langle\delta n^{2}\rangle\approx\biggl{(}1-\frac{\ell}{N+k}\biggr{)}\frac{\ell}{|s^{\prime\prime}(\lambda)|}. (131)

Altogether, the number ^c(L)\hat{\ell}_{c}(L) is given by the minimum \ell that satisfies

Cini+2s(λ)+2c(1N+k)|s′′(λ)||s(λ)|>k.C_{\rm ini}+2\ell s(\lambda)+2c\sqrt{\biggl{(}1-\frac{\ell}{N+k}\biggr{)}\frac{\ell}{|s^{\prime\prime}(\lambda)|}}|s^{\prime}(\lambda)|>k. (132)

By solving this, we obtain an explicit form of ^c(L)\hat{\ell}_{c}(L), which is

^c(L)0(L)+cfl(L),\hat{\ell}_{c}(L)\approx\ell_{0}(L)+c\ \!\ell_{\rm fl}(L), (133)

where

0(L)=Cini2s(λ)+k2s(λ),\displaystyle\ell_{0}(L)=-\frac{C_{\rm ini}}{2s(\lambda)}+\frac{k}{2s(\lambda)}, (134)
fl(L)=|s(λ)|s(λ)0(L)|s′′(λ)|(10(L)N+k).\displaystyle\ell_{\rm fl}(L)=\frac{|s\textquoteright(\lambda)|}{s(\lambda)}\sqrt{\frac{\ell_{0}(L)}{|s^{\prime\prime}(\lambda)|}\biggl{(}1-\frac{\ell_{0}(L)}{N+k}\biggr{)}}. (135)
Refer to caption
Figure 4: Numerical comparison of the delays computed in two ways. One is δΔ(λ):=Δ(λN,0)Δ(0,0)\delta\ell_{\Delta}(\lambda):=\ell_{\Delta}(\lambda N,0)-\ell_{\Delta}(0,0), based on the partial decoupling (Eq. (107)), and the other is δ^c(λ):=^c(λN)^c(0)\delta\hat{\ell}_{c}(\lambda):=\hat{\ell}_{c}(\lambda N)-\hat{\ell}_{c}(0) obtained from the clipping of entanglement (Eq. (133)). They are both plotted as functions of the ZZ-axis AM ratio λ=L/(N)\lambda=L/(\hbar N) of the initial Kerr BH BinB_{\rm in}. The upper and lower figures show the delays for the pure and mixed Kerr BHs, respectively. The size NN of the initial Kerr BH BinB_{\rm in} is set to 300300, and the size kk of the quantum information source AA is fixed to 33. In both figures, δΔ(λ)\delta\ell_{\Delta}(\lambda) is computed for Δ=0.005\Delta=0.005 (red plots), 0.050.05 (blue plots), and 0.50.5 (green plots), while δ^c(λ)\delta\hat{\ell}_{c}(\lambda) is computed for c=3.4c=3.4 (red dashed line), 2.62.6 (blue dashed line), and 1.61.6 (green dashed line) for the pure BH, and 10.810.8 (red dashed line), 8.78.7 (blue dashed line), and 6.26.2 (green dashed line) for the mixed BH. In both pure and mixed Kerr BHs, they nearly coincide, indicating that the argument of entanglement clipping provides good estimations.

In Fig. 4, we numerically compare ^c(L)\hat{\ell}_{c}(L) and Δ(L,0)\ell_{\Delta}(L,0). We observe that, by choosing proper cc and Δ\Delta, which both play the role of a smoothing parameter, they coincide very well. Hence, we conclude that the clipping argument provides a good estimate of the necessary number of qubits in the new radiation SradS_{\rm rad} for the information to become recoverable from the radiations.

In Eq. (133), we have decomposed ^c(L)\hat{\ell}_{c}(L) into two terms. The first term, 0(L)\ell_{0}(L), represents a delay originated from an information theoretic reason: when the number of qubits in SradS_{\rm rad} is increased by one, the degree of clipping is increased by 2s(λ)2s(\lambda) since a qubit has entropy s(λ)\approx s(\lambda). Thus, |Srad||S_{\rm rad}| should be at least 0(L)\ell_{0}(L) so as to cancel the initial negative degree of clipping CiniC_{\rm ini}, and more for SradS_{\rm rad} to typically have a sufficiently large space to store 2k2^{k} states without significant overlaps. That is why we refer to 0(L)\ell_{0}(L) as an information-theoretic delay.

It is the second term fl(L)\ell_{\rm fl}(L) that is a non-trivial consequence of the symmetry of the Kerr BH. It stems from the fluctuation of the ZZ-axis AM in the radiation, making the clipping condition harder to be fulfilled. We hence call fl(L)\ell_{\rm fl}(L) a fluctuational delay. Note also that, when the ZZ-axis AM LL of the initial Kerr BH is small, the delay of information leakage, δ^c(L):=^c(L)^c(0)\delta\hat{\ell}_{c}(L):=\hat{\ell}_{c}(L)-\hat{\ell}_{c}(0), satisfies

δ^c(L)fl(L),\delta\hat{\ell}_{c}(L)\propto\ell_{\rm fl}(L), (136)

since 0(L)0(0)\ell_{0}(L)\approx\ell_{0}(0) and fl(0)=0\ell_{\rm fl}(0)=0 for small LL.

6.1.3 Delay and BH Thermodynamics

The fluctuational delay fl(L)\ell_{\rm fl}(L) can be further rewritten in terms of thermodynamic quantities of the Kerr BH. In this context, it is more natural to quantify the delay by the amount of the ZZ-axis AM rather than the number of qubits. We hence introduce LflL_{\rm fl} by

Lfl:=λfl(L),L_{\rm fl}:=\hbar\lambda\ell_{\rm fl}(L), (137)

and investigate it. In this Subsection, we explicitly write the Plank constant \hbar. Correspondingly, we also introduce L0:=λ0(L)L_{0}:=\hbar\lambda\ell_{0}(L), which is the ZZ-axis AM corresponding to the information-theoretic delay 0(L)\ell_{0}(L).

In order to consider the thermodynamics of the Kerr BH, it is necessary to take its Hamiltonian into account. The Hamiltonian should be invariant under the rotation around the ZZ-axis and also sufficiently random for the dynamics to be symmetry-preserving scrambling. Since we are interested in the thermodynamics properties that do not strongly rely on the randomness, we especially consider the Kerr BH with temperature TT sufficiently higher than the energy scale of the Hamiltonian, so that the thermodynamic entropy S(λ,L,T)S(\lambda,L,T) of the Kerr BH is independent of TT, and the free energy F(T,λ,L)F(T,\lambda,L) is approximately given by TS(λ,L)-TS(\lambda,L).

Let ω:=(FL)T,N\omega:=-\bigl{(}\frac{\partial F}{\partial L}\bigl{)}_{T,N} be an intensive state function conjugate to the ZZ-axis AM LL, and α:=(1LLT)ω,N\alpha:=\bigl{(}\frac{1}{L}\frac{\partial L}{\partial T}\bigl{)}_{\omega,N} be the sensitivity of the ZZ-axis AM to the temperature. More specifically, the state function ω(T,λ)\omega(T,\lambda) is the angular velocity of the Kerr BH. Since the thermodynamic entropy S(λ,L)S(\lambda,L) is given by

S(λ,L)=kBs(λ)N,S(\lambda,L)=k_{B}s(\lambda)N, (138)

where kBk_{B} is the Boltzmann constant, we have

ω(T,λ)=kBTs(λ),\displaystyle\omega(T,\lambda)=\frac{k_{B}Ts^{\prime}(\lambda)}{\hbar}, (139)
α(T,λ)=s(λ)s′′(λ)λT,\displaystyle\alpha(T,\lambda)=-\frac{s^{\prime}(\lambda)}{s^{\prime\prime}(\lambda)\lambda T}, (140)

in our particular model of the Kerr BH. Hence, it follows that

|s(λ)2s′′(λ)|=λkB|ω(T,L)α(T,L)|.\biggl{|}\frac{s^{\prime}(\lambda)^{2}}{s^{\prime\prime}(\lambda)}\biggr{|}=\frac{\lambda\hbar}{k_{B}}|\omega(T,L)\alpha(T,L)|. (141)

Using these and assuming kNk\ll N, we obtain from Eqs. (135) and (137) that

LflLS(λ,L)kB|ω(T,λ)α(T,λ)|(1L0L)L0.\displaystyle L_{\rm fl}\approx\frac{L}{S(\lambda,L)}\sqrt{k_{B}\bigl{|}\omega(T,\lambda)\alpha(T,\lambda)\bigr{|}\biggl{(}1-\frac{L_{0}}{L}\biggr{)}L_{0}}. (142)

Note that ω(T,λ)α(T,λ)\omega(T,\lambda)\alpha(T,\lambda) is independent of the temperature TT, and so is LflL_{\rm fl}.

After radiating the ZZ-axis AM by L0L_{0} to fulfill the information-theoretic requirement, the Kerr BH must further radiate an extra amount LflL_{\rm fl} of the ZZ-axis AM to release the information. It is clear from Eq. (142) that this amount LflL_{\rm fl} is solely determined by intensive thermodynamic quantities of the initial Kerr BH, that is, the ZZ-axis AM per entropy L/S(λ,L)L/S(\lambda,L), the angular velocity ω(T,λ)\omega(T,\lambda) of the Kerr BH, and the coefficient α(T,λ)\alpha(T,\lambda) of the thermal sensitivity of the ZZ-axis AM.

The delay of information leakage is, thus, closely related to the thermodynamic properties of the BH. This, in turn, implies that, if we understand the thermodynamics of the BH well, it is possible to predict how quickly the BH releases the quantum information therein.

We emphasize that the above argument is likely to be applicable to the system with any global abelian symmetry. A particularly important application is the case where energy is conserved during the unitary dynamics of the system. Assuming that the dynamics is energy-preserving scrambling in the sense that it scrambles only quantum states with the same energy scale, we can evaluate the delay of information leakage based on Eq. (135). Suppose that the initial quantum many-body system BH is thermodynamical and has energy EE with negligibly small fluctuation. The thermodynamic entropy S(E)S(E) of the system, where EE is the internal energy, and the heat capacity CVC_{V} satisfy

|CV|=kB|S(E)2S′′(E)|.|C_{V}|=k_{B}\biggl{|}\frac{S^{\prime}(E)^{2}}{S^{\prime\prime}(E)}\biggr{|}. (143)

Thus, a similar calculation from Eq. (135) leads to

EflES(E)kB|CV|(1E0E)E0E,E_{\rm fl}\approx\frac{E}{S(E)}\sqrt{k_{B}|C_{V}|\biggl{(}1-\frac{E_{0}}{E}\biggr{)}\frac{E_{0}}{E}}, (144)

where E0E_{0} and EflE_{\rm fl} are the energy corresponding to the information-theoretic and fluctuational delays, respectively. As E0/EE_{0}/E is between 0 and 11, the amount is basically determined by the ratio between the energy EE and entropy S(E)S(E) of the initial black hole, and the heat capacity CVC_{V}, connecting the information leakage to thermodynamic quantities of the quantum many-body system.

6.2 Information remnant and symmetry-breaking

We finally investigate the origin of the information remnant characterized by ηξ\eta_{\xi} (see Eq. (85)). As explained in Subsec. 5.5, an important feature of ηξ\eta_{\xi} is that it depends on the initial fluctuation δL\delta L of the ZZ-axis AM in BinB_{\rm in} but not on the ZZ-axis AM LL itself.

We start with a simple observation that the information remnant should exist when the unitary dynamics of the BH is symmetric for the following reason. First, throwing the system AA into the initial BH BinB_{\rm in} changes the ZZ-axis AM of the BH, which remains unchanged by the subsequent unitary dynamics in S=ABinS=AB_{\rm in}. When the system SS is eventually split to two random subsystems SinS_{\rm in} and SradS_{\rm rad}, the change induced by throwing AA into the BH is inherited to both SinS_{\rm in} and SradS_{\rm rad}. This implies that the original value of the ZZ-axis AM in AA, which is kept stored in the reference system RR, can be inferred to some extent from the remaining BH SinS_{\rm in} by measuring its ZZ-axis AM. Hence, the information in AA about the coherence between difference values of the ZZ-axis AM cannot be fully accessed from the radiation SradS_{\rm rad}.

This observation also implies that the amount of information remnant should be related to that of the fluctuation of the ZZ-axis AM in the remaining BH SinS_{\rm in}: if the fluctuation overwhelms the change in the ZZ-axis AM caused by throwing AA into the BH, the information of AA shall not be obtained from the AM in the BH SinS_{\rm in}, making the information remnant negligible. To evaluate this, we observe that, after SS is split into SinS_{\rm in} and SradS_{\rm rad} of N+kN+k-\ell and \ell qubits, respectively, the initial change caused by throwing of AA into the BH is also split. Since the system is split randomly, the change inherited to SinS_{\rm in} is approximately given by

N+kN+kO(k)=(1N+k)O(k)\frac{N+k-\ell}{N+k}O(k)=\biggl{(}1-\frac{\ell}{N+k-\ell}\biggr{)}O(k) (145)

on average. If this change is much smaller than the fluctuation in SinS_{\rm in}, then the information remnant is expected to be negligible. For instance, this is the case for any \ell when the initial fluctuation δL\delta L in BinB_{\rm in} satisfies δLk\delta L\gg k. In contrast, if δLk\delta L\ll k, the information remnant remains non-negligible until sufficiently large amount of radiation is emitted such that the fluctuation in SinS_{\rm in} becomes much larger than (1N+k)k\bigl{(}1-\frac{\ell}{N+k-\ell}\bigr{)}k. As the \ell-qubit evaporation increases the fluctuation of the BH by O()O(\sqrt{\ell}), non-negligible information remnant must exist if (1N+k)k\bigl{(}1-\frac{\ell}{N+k-\ell}\bigr{)}k\leq\sqrt{\ell} as order estimation.

In the following, we make this argument rigorous and provide a quantitative estimate of the amount of information remnant. Since we are interested in the information remnant, we consider only the situation where the new radiation SradS_{\rm rad} is sufficiently large so that the Kerr BH has already get partially decoupled and the state therein and the reference RR is ΓSinR\Gamma^{S_{\rm in}R}. The state is explicitly given by (see Proposition 3)

ΓSinR=m=0N+k2N+k(N+km)TrS[ΠmS(πSradΦSinSin)]TrS[ΠmS(ΦARξBin)].\Gamma^{S_{\rm in}R}=\sum_{m=0}^{N+k}\frac{2^{N+k}}{\binom{N+k}{m}}\operatorname{Tr}_{S^{\prime}}[\Pi_{m}^{S^{\prime}}(\pi^{S^{\prime}_{\rm rad}}\otimes\Phi^{S_{\rm in}S^{\prime}_{\rm in}})]\\ \otimes\operatorname{Tr}_{S}[\Pi_{m}^{S}(\Phi^{AR}\otimes\xi^{B_{\rm in}})]. (146)

Let us now consider to what extent SinS_{\rm in} has the information of the ZZ-axis AM originally in AA. Since the value of the ZZ-axis AM in AA is stored in the reference RR, this can be checked by measuring the ZZ-axis AM in RR and by evaluating how much one can estimate the measurement outcome of RR from SinS_{\rm in}. There should be various ways of estimating the outcome in RR from SinS_{\rm in}, but we particularly consider an estimation by measuring the ZZ-axis AM in SinS_{\rm in}. This provides a lower bound of the information remnant.

This situation is formulated by the projective measurement {ΠκR}\{\Pi_{\kappa}^{R}\} on RR for its ZZ-axis AM and that {ΠνSin}\{\Pi_{\nu}^{S_{\rm in}}\} on SinS_{\rm in}. The labels κ\kappa and ν\nu represent each ZZ-axis AM, and Πκ\Pi_{\kappa} and Πν\Pi_{\nu} are the projectors on the corresponding subspaces. When the measured system SinRS_{\rm in}R is in state ΓSinR\Gamma^{S_{\rm in}R}, the marginal probabilities to obtain ν\nu in SinS_{\rm in} and κ\kappa in RR are, respectively, given by

q(κ)\displaystyle q(\kappa) :=Tr[ΠκRΓR]=Tr[ΠκRπR],\displaystyle:=\operatorname{Tr}[\Pi_{\kappa}^{R}\Gamma^{R}]=\operatorname{Tr}[\Pi_{\kappa}^{R}\pi^{R}], (147)
P(ν)\displaystyle P(\nu) :=Tr[ΠνSinΓSin],\displaystyle:=\operatorname{Tr}[\Pi_{\nu}^{S_{\rm in}}\Gamma^{S_{\rm in}}], (148)

where ΓκSin:=TrR[ΠκRΓSinR]/q(κ)\Gamma^{S_{\rm in}}_{\kappa}:=\operatorname{Tr}_{R}[\Pi_{\kappa}^{R}\Gamma^{S_{\rm in}R}]/q(\kappa). We also use the conditional probability

P(ν|κ)\displaystyle P(\nu|\kappa) :=Tr[(ΠνSinΠκR)ΓSinR]/q(κ),\displaystyle:=\operatorname{Tr}\bigl{[}(\Pi_{\nu}^{S_{\rm in}}\otimes\Pi_{\kappa}^{R})\Gamma^{S_{\rm in}R}]/q(\kappa), (149)
=Tr[ΠνSinΓκSin],\displaystyle=\operatorname{Tr}\bigl{[}\Pi_{\nu}^{S_{\rm in}}\Gamma^{S_{\rm in}}_{\kappa}], (150)

which satisfies

P(ν)=κq(κ)P(ν|κ).P(\nu)=\sum_{\kappa}q(\kappa)P(\nu|\kappa). (151)

Using these probabilities, ηξ=ΓSinRΓSinπR1\eta_{\xi}=\|\Gamma^{S_{\rm in}R}-\Gamma^{S_{\rm in}}\otimes\pi^{R}\|_{1} (see Eq. (85)) is bounded from below by

ηξκq(κ)ν|P(ν|κ)P(ν)|.\eta_{\xi}\geq\sum_{\kappa}q(\kappa)\sum_{\nu}\bigl{|}P(\nu|\kappa)-P(\nu)\bigr{|}. (152)

This follows from the monotonicity of the trace distance under the CPTP map for transformation ρSinRν,κTr[(ΠνSinΠκR)ρSinR]|ν,κν,κ|\rho^{S_{\rm in}R}\mapsto\sum_{\nu,\kappa}\operatorname{Tr}[(\Pi_{\nu}^{S_{\rm in}}\otimes\Pi_{\kappa}^{R})\rho^{S_{\rm in}R}]|\nu,\kappa\rangle\langle\nu,\kappa| with |ν,κ|\nu,\kappa\rangle being the mutually orthonormal states in an ancillary system.

For the sake of simplicity of the analysis, we now approximate the value of ν\nu by continuous values and replace the above probabilities P(ν)P(\nu) and P(ν|κ)P(\nu|\kappa) with probability density functions p(ν)p(\nu) and p(ν|κ)p(\nu|\kappa), respectively. We also assume that the κ\kappa-dependence of p(ν|κ)p(\nu|\kappa) is approximated by a shift without changing its common shape, which is given by a probability function p¯(ν)\bar{p}(\nu) with vanishing tails for a large deviation |νν||\nu-\langle\nu\rangle|. More specifically, we assume that

p(ν|κ)p¯(ναδκ),p(\nu|\kappa)\approx\bar{p}(\nu-\alpha\delta\kappa), (153)

where p¯(μ):=p(μ|κ=κ)\bar{p}(\mu):=p(\mu|\kappa=\langle\kappa\rangle) with α=1/(N+k)\alpha=1-\ell/(N+k), δκ=κκ\delta\kappa=\kappa-\langle\kappa\rangle, and \langle\cdot\rangle is the expectation over the probability distribution q(κ)q(\kappa). Note that the rescaling by α\alpha in Eq. (153) is needed since SS is composed of N+kN+k qubits while SinS_{\rm in} is of N+kN+k-\ell qubits. When |δκ||\delta\kappa| is sufficiently small, we have

p¯(ναδκ)=p¯(ν)αδκdp¯(ν)dν+O(δκ2),\bar{p}(\nu-\alpha\delta\kappa)=\bar{p}(\nu)-\alpha\delta\kappa\frac{d\bar{p}(\nu)}{d\nu}+O(\delta\kappa^{2}), (154)

which further implies that p(ν)=p¯(ν)+O(δκ2)p(\nu)=\bar{p}(\nu)+O(\delta\kappa^{2}) due to Eq. (151) and δκ=0\langle\delta\kappa\rangle=0.

Refer to caption
Figure 5: Figure (i) explains the ZZ-axial symmetry of a pure state and the Q function. For a pure state |ξBin|\xi\rangle^{B_{\rm in}}, the Q function is defined with variables LXL_{X} and LZL_{Z}, which correspond to the XX- and ZZ-axis AMs, respectively, by Qξ((N/2)1/2LX,(N/2)1/2LZ)=|θ,ϕ|ξBin|2Q_{\xi}\bigl{(}(N/2)^{-1/2}L_{X},(N/2)^{-1/2}L_{Z}\bigr{)}=\bigl{|}\langle\theta,\phi|\xi\rangle^{B_{\rm in}}\bigr{|}^{2}, where |θ,ϕ:=(cosθ2|0+eiϕsinθ2|1)N|\theta,\phi\rangle:=(\cos\frac{\theta}{2}|0\rangle+e^{i\phi}\sin\frac{\theta}{2}|1\rangle)^{\otimes N} (|ϕ|π/2|\phi|\leq\pi/2) and (θ,ϕ)(\theta,\phi) are those satisfying sinθcosϕ=2LX/N\sin\theta\cos\phi=2L_{X}/N and cosθ=2LZ/N\cos\theta=2L_{Z}/N. Note that the function is regarded as the Husimi Q function for NN\rightarrow\infty when LYN/2\langle L_{Y}\rangle\approx N/2. In Figure (i), we depict |θ,ϕ|ξBin|2|\langle\theta,\phi|\xi\rangle^{B_{\rm in}}|^{2}: since it is a function of (θ,ϕ)(\theta,\phi), it can be visualized on a sphere. The Q function is then depicted by projecting the sphere surface on XZXZ-plane, as shown in the right side of Figure (i) for two pure states as examples. The states |ξ|\xi\rangle are taken in the form of Eq. (98) for δL=0\delta L=0 and δL=0.9N\delta L=0.9\sqrt{N}, where |φn|\varphi_{n}\rangle is especially chosen as the state of equal superposition of all product basis states in nBin{\mathcal{H}}_{n}^{B_{\rm in}}. The state for δL=0\delta L=0 is apparently invariant under the rotation around the ZZ-axis: since the number of up spins is constant, the rotation changes only the global phase. This is visualized by the Q function, which is independent of LXL_{X}. This is not the cases for δL=0.9N\delta L=0.9\sqrt{N}: the state is variant and the figure is non-symmetric if they are rotated around the ZZ-axis. These indicate that the Q function visualizes how symmetric the states are. Figure (ii) shows the semi-log plot of ηξ(N,k,)\eta_{\xi}(N,k,\ell) for various pure states ξ\xi of the initial Kerr BH BinB_{\rm in} with different fluctuations of the ZZ-axis AM, as well as the corresponding Q functions. The fluctuation δL\delta L is taken as δL=0\delta L=0 (black), 0.1N0.1\sqrt{N} (red), 0.5N0.5\sqrt{N} (blue), 0.9N0.9\sqrt{N} (green), and 0.3N0.3N (brown). The size NN of the initial Kerr BH is fixed to 10001000, and kk is set to 11. Note that the average ZZ-axis AM LL can be arbitrary since ηξ(N,k,)\eta_{\xi}(N,k,\ell) does not depend on LL. Comparing this with the Q functions for each ξ(L,δL)\xi(L,\delta L), we observe that ηξ(N,k,)\eta_{\xi}(N,k,\ell) is small when the degree of symmetry-breaking is large.

Using these approximations and replacing the summation over ν\nu in Eq. (152) by integral, we obtain

ηξ\displaystyle\eta_{\xi} ακq(κ)|δκ|dν|dp¯(ν)dν|+O(δκ2),\displaystyle\geq\alpha\sum_{\kappa}q(\kappa)|\delta\kappa|\int d\nu\bigl{|}\frac{d\bar{p}(\nu)}{d\nu}\bigr{|}+O(\delta\kappa^{2}), (155)
=α|δκ|dν|dp¯(ν)dν|+O(δκ2)\displaystyle=\alpha\langle|\delta\kappa|\rangle\int d\nu\biggl{|}\frac{d\bar{p}(\nu)}{d\nu}\biggr{|}+O(\delta\kappa^{2}) (156)

where |δκ|:=κq(κ)|δκ|\langle|\delta\kappa|\rangle:=\sum_{\kappa}q(\kappa)|\delta\kappa| is the mean absolute deviation of the ZZ-axis AM in RR. Recalling that we have assumed vanishing tails for p¯(ν)\bar{p}(\nu), the integral in Eq. (156) can be bounded from below by 2maxνp¯(ν)2\max_{\nu}\bar{p}(\nu). Thus, we have

ηξ\displaystyle\eta_{\xi} 2α|δκ|maxνp¯(ν)+O(δκ2).\displaystyle\geq 2\alpha\langle|\delta\kappa|\rangle\max_{\nu}\bar{p}(\nu)+O(\delta\kappa^{2}). (157)

The value maxνp¯(ν)\max_{\nu}\bar{p}(\nu) can be further bounded from below in terms of the variance δν2\langle\delta\nu^{2}\rangle of the ZZ-axis AM in SinS_{\rm in}. This follows from the fact that the probability distribution that has the least standard deviation under the condition that the maximum probability is given is the rectangle function r(x)=rmaxr(x)=r_{\rm max} for x[1/(2rmax),1/(2rmax)]x\in[-1/(2r_{\rm max}),1/(2r_{\rm max})] and 0 otherwise. In that case, the variance VV is

V=1/(2rmax)1/(2rmax)x2rmax𝑑x=112rmax2.V=\int_{-1/(2r_{\rm max})}^{1/(2r_{\rm max})}x^{2}r_{\rm max}dx=\frac{1}{12\ \!r_{\rm max}^{2}}. (158)

Hence, for any probability density function p(x)p(x), it holds that maxp(x)(23V)1\max p(x)\geq(2\sqrt{3}\sqrt{V})^{-1}. Applying this to our case, we obtain maxνp¯(ν)(23δν2)1\max_{\nu}\bar{p}(\nu)\geq(2\sqrt{3}\sqrt{\langle\delta\nu^{2}\rangle})^{-1}. We thus arrive at

ηξ13|δκ|(1N+k)1δν2+O(δκ2).\eta_{\xi}\geq\frac{1}{\sqrt{3}}\langle|\delta\kappa|\rangle\biggl{(}1-\frac{\ell}{N+k}\biggr{)}\frac{1}{\sqrt{\langle\delta\nu^{2}\rangle}}+O(\delta\kappa^{2}). (159)

Since δν2\langle\delta\nu^{2}\rangle is the variance of the ZZ-axis AM in the BH SinS_{\rm in}, this confirms the expectation that the information remnant is bounded from below by the fluctuation of the AM in the BH. In particular, when the fluctuation is small, the amount of information remnant should be necessarily large as expected.

We continue our analysis and show that the fluctuation is related to how much symmetric the BH state is. Due to Eq. (159), this further implies that the information remnant is also characterized by the symmetry of the state of the BH. To start with, in Fig. 5, the degree of symmetry of the state ξBin\xi^{B_{\rm in}} is visualized by using the so-called Q function. There, we consider only pure states for demonstration. By comparing the Q function of the state with the corresponding ηξ\eta_{\xi}, it is clear that the ZZ-axial symmetry in ξBin\xi^{B_{\rm in}} is strongly broken when ηξ\eta_{\xi} is small.

To analyze this relation in a quantitative manner, we introduce the degree of symmetry breaking by the sensitivity of the state under the symmetric action. Let |ΨSinSin|\Psi\rangle^{S_{\rm in}S^{\prime}_{\rm in}} be a purification of the state ΨSin\Psi^{S_{\rm in}}, and Bθ(Ψ)B_{\theta}(\Psi) denote

Bθ(Ψ):=1|Ψ|eiθLZI|Ψ|2,B_{\theta}(\Psi):=1-\bigl{|}\langle\Psi|e^{i\theta L_{Z}}\otimes I|\Psi\rangle\bigr{|}^{2}, (160)

where LZL_{Z} is the ZZ-axis AM operator in SinS_{\rm in} and the second identity acts on SinS^{\prime}_{\rm in}. Then, we define the degree ζ(Sin)\zeta(S_{\rm in}) of symmetry-breaking by

ζ(Sin):=2θ2Bθ(Ψ)|θ=0.\zeta(S_{\rm in}):=\frac{\partial^{2}}{\partial\theta^{2}}B_{\theta}(\Psi)\biggl{|}_{\theta=0}. (161)

This quantifies how sensitive the state Ψ\Psi is to the symmetric action, which is just a rotation around the ZZ-axis in our case. Note that we here considered SinS_{\rm in}, but the degree of symmetry-breaking can be similarly defined in any system, such as in the initial BH BinB_{\rm in}.

The degree ζ\zeta of symmetry-breaking is closely related to the variance δν2\langle\delta\nu^{2}\rangle of the ZZ-axis AM ν\nu in ΨSin\Psi^{S_{\rm in}}. Since we have

|Ψ|eiθLZI|Ψ|2\displaystyle\bigl{|}\langle\Psi|e^{i\theta L_{Z}}\otimes I|\Psi\rangle\bigr{|}^{2}
=|Tr[eiθLZΨ]|2\displaystyle=\bigl{|}\operatorname{Tr}[e^{i\theta L_{Z}}\Psi]\bigr{|}^{2} (162)
=|1+iθTr[LZΨ]θ22Tr[LZ2Ψ]|2+O(θ3)\displaystyle=\bigl{|}1+i\theta\operatorname{Tr}[L_{Z}\Psi]-\frac{\theta^{2}}{2}\operatorname{Tr}[L_{Z}^{2}\Psi]\bigr{|}^{2}+O(\theta^{3}) (163)
=1+θ2((Tr[LZΨ])2Tr[LZ2Ψ])+O(θ3)\displaystyle=1+\theta^{2}\bigl{(}(\operatorname{Tr}[L_{Z}\Psi])^{2}-\operatorname{Tr}[L_{Z}^{2}\Psi]\bigr{)}+O(\theta^{3}) (164)
=1θ2δν2+O(θ3),\displaystyle=1-\theta^{2}\langle\delta\nu^{2}\rangle+O(\theta^{3}), (165)

we obtain ζ(Sin)=2δν2\zeta(S_{\rm in})=2\langle\delta\nu^{2}\rangle.

Substituting this relation between the variance and the degree of symmetry-breaking into Eq. (159), it follows that

ηξ23|δκ|(1N+k)1ζ(Sin).\eta_{\xi}\geq\sqrt{\frac{2}{3}}\langle|\delta\kappa|\rangle\biggl{(}1-\frac{\ell}{N+k}\biggr{)}\frac{1}{\sqrt{\zeta(S_{\rm in})}}. (166)

Thus, when information remnant is small, ζ(Sin)\zeta(S_{\rm in}) should be necessarily large, implying that the ZZ-axial symmetry in the remaining Kerr BH SinS_{\rm in} should be broken strongly. This clarifies yet another micro-macro correspondence of a quantum Kerr BH.

The symmetry-breaking in SinS_{\rm in} has two origins: one is that in the initial Kerr BH BinB_{\rm in}, and the other is due to the new radiation, where \ell-qubit radiation increases the standard deviation of the ZZ-axis AM in the remaining Kerr BH SinS_{\rm in} by O()O(\sqrt{\ell}), and so does the degree of symmetry breaking. When δL\sqrt{\ell}\ll\delta L with δL\delta L being the standard deviation of the ZZ-axis AM in the initial Kerr BH BinB_{\rm in}, the former origin is dominant. On the other hand, when δL\sqrt{\ell}\gg\delta L, the symmetry-breaking in SinS_{\rm in} is mostly originated from the radiation process. We thus conclude that

ηξ{a21δLwhen δL,awhen δL,\eta_{\xi}\gtrsim\begin{cases}\frac{a}{\sqrt{2}}\frac{1}{\delta L}&\text{when $\sqrt{\ell}\ll\delta L$},\\ \frac{a}{\sqrt{\ell}}&\text{when $\sqrt{\ell}\gg\delta L$},\end{cases} (167)

where a=2/3|δκ|(1/(N+k))a=\sqrt{2/3}\langle|\delta\kappa|\rangle(1-\ell/(N+k)). Here, we have used the fact that the degree of symmetry breaking in BinB_{\rm in}, ζ(Bin)\zeta(B_{\rm in}), is given by 2(δL)22(\delta L)^{2}.

In the case of the pure Kerr BH, the information remnant is negligible in the thermodynamic limit (NN\rightarrow\infty), assuming that kk remains constant. This is because the partial decoupling occurs in the pure Kerr BH only when N/2\ell\gtrsim N/2. Thus, Eq. (167) implies that, regardless of δL\delta L, the information remnant vanishes when NN\rightarrow\infty. In contrast, for the mixed Kerr BH, there are cases where the partial decoupling sets in when =O(k)\ell=O(k). In this case, the fluctuation δL\delta L of the ZZ-axis AM in the initial Kerr BH BinB_{\rm in} determines the information remnant. In particular, if δL\delta L is not so large that the lower case of Eq. (167) holds, a non-negligible amount of information remains in the Kerr BH even in the thermodynamic limit unless \ell becomes sufficiently large. This is in sharp contrast to the BH without symmetry, in which the information of kk qubits can be almost fully recovered from =O(k)\ell=O(k) radiation. Thus, the information remnant is a substantial feature induced by the symmetry when the initial Kerr BH is sufficiently mixed with small fluctuation δL\delta L of the ZZ-axis AM.

Finally, we refer to the recent result [54], where a rigorous lower bound on the information remnant for any symmetry was derived in terms of the quantum Fisher information of the initial/remaining BH. Since the quantum Fisher information is another characterization of variance, our and their results are consistent with each other.

7 Conclusion and Discussions

In this paper, we have investigated the Hayden-Preskill protocol when the system conserves the number of up-spins, or equivalently the ZZ-component of the angular momentum. Based on the partial decoupling approach, we have first provided general formulas for upper bounds on the errors in recovering the information. From the numerical evaluations of the recovery errors, we have shown that the symmetry induces two substantial deviations from the protocol without symmetry. One is the delay of information leakage and the other is the information remnant. Depending on the initial condition of the system, these phenomena can be negligible or macroscopically large. This is of particular interest since it implies that there are cases where symmetry substantially changes the process of information leakage from that without symmetry.

We have then investigated the delay and the information remnant from the macroscopic physics point of view. By introducing the clipping of entanglement, we have revealed that the delay is characterized by the thermodynamic properties of the initial system. The close relation between the information remnant and the symmetry-breaking of the system has been also revealed. These relations establish connections between the information leakage problem and macroscopic physics.

The results obtained by partial decoupling can be easily generalized to the systems with any abelian global symmetries, such as energy conservation. It is also possible to apply the partial decoupling approach to the systems with non-abelian global symmetry. See Ref. [47] for the more general form of partial decoupling. We, however, leave the non-abelian case as an interesting future problem. In that case, the information cannot be simply divided into the symmetry-invariant and the other ones, and more careful analysis will be needed.

Another important open problem is about an approximate realization of the symmetry-preserving Haar scrambling. All of our analyses rely on the assumption that the internal unitary dynamics of the system is symmetry-preserving scrambling, which will not be rigorously satisfied since it takes exponentially long time to implement a Haar random unitary. Hence, it is of crucial importance to consider approximate implementations of symmetry-preserving scrambling. Approximations of random unitary have been studied in terms of unitary designs, and a couple of results are known about the time needed to achieve unitary designs [55, 56, 57, 58]. It will be of great interest to investigate how symmetries affect such results [15, 59]. More explicitly, it is interesting to address if we can construct symmetry-preserving unitary designs by quantum circuits.

It is also of interest to explicitly construct a decoder for the Hayden-Preskill protocol in the presence of symmetry. Without symmetry, a couple of decoders were proposed in the literature [60, 61]. It is not immediately clear if one can construct a decoder in the same way, which will be an interesting open problem.

8 Acknowledgements

Y.N. is supported by by JST, PRESTO Grant Number JPMJPR1865, Japan, and partially by JST CREST, Grant Number JPMJCR1671, Japan. Y.N. is partially supported by Grants-in-Aid for Transformative Research Areas (A) No. JP21H05182 and No. JP21H05183 from MEXT of Japan, as well as by JSPS KAKENHI Grant Number JP22K03464, Japan. Y.N. and M.K. are supported by JST Moonshot R&D, Grant Number JPMJMS2061, Japan.

References

Appendix A Proof of Theorem 1

To prove Theorem 1, we use the theorem shown in Ref. [47] and the concentration of measure phenomena for product measures.

Theorem 5 (Non-smoothed partial decoupling theorem [47]).

In the situation described in Subsec. 5.1, it holds that

𝔼US𝖧×||SE(USϱSRUS)ΓER||1212Hmin(S|ER)Γ,\mathbb{E}_{U^{S}\sim{\sf H}_{\times}}\bigl{|}\!\bigl{|}\mathcal{M}^{S\rightarrow E}\bigl{(}U^{S}\varrho^{SR}U^{S\dagger}\bigr{)}-\Gamma^{ER}\bigr{|}\!\bigr{|}_{1}\\ \leq 2^{-\frac{1}{2}H_{\rm min}(S^{*}|ER)_{\Gamma}}, (168)

where Hmin(S|ER)ΓH_{\rm min}(S^{*}|ER)_{\Gamma} is the conditional min-entropy of ΓSER\Gamma^{S^{*}ER}.

It is well-known that a Haar random unitary has concentration properties when the degree of the group is large [62], which finds a number of applications in quantum information science. In most cases, the Haar measure on the whole unitary group is considered. However, the concentration also happens for the product measure 𝖧×=𝖧1××𝖧J{\sf H}_{\times}={\sf H}_{1}\times\cdots\times{\sf H}_{J}.

To explain the concentration of measure on the product space, let us introduce the L2L_{2}-sum of the Hilbert-Schmidt norms. Let 𝖴×{\sf U}_{\times} be a product of unitary groups 𝖴(d1)××𝖴(dJ){\sf U}(d_{1})\times\dots\times{\sf U}(d_{J}). For U=(U1,,UJ)𝖴×U=(U_{1},\cdots,U_{J})\in{\sf U}_{\times} and V=(V1,,VJ)𝖴×V=(V_{1},\cdots,V_{J})\in{\sf U}_{\times}, the L2L_{2}-sum D(U,V)D(U,V) of the Hilbert-Schmidt norms on 𝖴×{\sf U}_{\times} is defined by

D(U,V)=j𝒥UjVj22.D(U,V)=\sqrt{\sum_{j\in{\mathcal{J}}}\|U_{j}-V_{j}\|_{2}^{2}}. (169)

We also use Lipschitz functions. A real-valued function FF on a metric space (X,d)(X,d) is said to be Lipschitz if

FLip:=supxyX|F(x)F(y)|d(x,y)<.\|F\|_{\rm Lip}:=\sup_{x\neq y\in X}\frac{|F(x)-F(y)|}{d(x,y)}<\infty. (170)

The quantity FLip\|F\|_{\rm Lip} is called an Lipschitz constant of FF. The function FF with a Lipschitz constant LL is called LL-Lipschitz.

The following provides the concentrating property on the product space.

Theorem 6 (Concentration of measure on the product space [63]).

Let 𝖴×=𝖴(d1)××𝖴(dJ){\sf U}_{\times}={\sf U}(d_{1})\times\dots\times{\sf U}(d_{J}) equipped with the L2L_{2}-sum of Hilbert-Schmidt norms, and 𝖧×=𝖧1××𝖧J{\sf H}_{\times}={\sf H}_{1}\times\cdots\times{\sf H}_{J} be the product probability measure. Suppose that a function F:𝖴×F:{\sf U}_{\times}\rightarrow\mathbb{R} is LL-Lipschitz. Then, for every δ>0\delta>0,

F(U1,,UJ)𝔼(U1,,UJ)𝖧×[F]+δ,F(U_{1},\dots,U_{J})\geq\mathbb{E}_{(U_{1},\cdots,U_{J})\sim{\sf H}_{\times}}[F]+\delta, (171)

with probability at most exp[δ2dmin12L2]\exp[-\frac{\delta^{2}d_{\rm min}}{12L^{2}}], where dmin=min{d1,,dJ}d_{\rm min}=\min\{d_{1},\dots,d_{J}\}.

Based on Theorems 5 and 6, we now prove Theorem 1 by identifying the unitary US=j𝒥UjSU^{S}=\bigoplus_{j\in{\mathcal{J}}}U^{S}_{j} as a point (U1S,,UJS)(U_{1}^{S},\dots,U_{J}^{S}) on 𝖴×=𝖴(d1)××𝖴(dJ){\sf U}_{\times}={\sf U}(d_{1})\times\cdots\times{\sf U}(d_{J}).

Proof (Theorem 1).

For a given US=j𝒥UjSU^{S}=\bigoplus_{j\in{\mathcal{J}}}U^{S}_{j}, let (U1S,,UJS)(U_{1}^{S},\dots,U_{J}^{S}) be the corresponding point on 𝖴×{\sf U}_{\times} equipped with the L2L_{2}-sum of Hilbert-Schmidt norms DD. By a slight abuse of notation, we also denote the point by USU^{S}.

Let F:𝖴×F:{\sf U}_{\times}\rightarrow\mathbb{R} be a function given by

F(U1S,,UJS):=SE(USϱSRUS)TrS[ΓSER]1.F(U_{1}^{S},\dots,U_{J}^{S})\\ :=\bigl{\|}{\mathcal{M}}^{S\rightarrow E}\bigl{(}U^{S}\varrho^{SR}U^{S\dagger}\bigr{)}-\operatorname{Tr}_{S^{*}}[\Gamma^{S^{*}ER}]\bigr{\|}_{1}. (172)

We compute the Lipschitz constant of FF. Using the triangle inequality, we obtain

|F(U1S,,UJS)F(V1S,,VJS)|\displaystyle|F(U_{1}^{S},\dots,U_{J}^{S})-F(V_{1}^{S},\dots,V_{J}^{S})|
SE(USΨSRUS)SE(VSϱSRVS)1\displaystyle\leq\|{\mathcal{M}}^{S\rightarrow E}(U^{S}\Psi^{SR}U^{S\dagger})-{\mathcal{M}}^{S\rightarrow E}(V^{S}\varrho^{SR}V^{S\dagger})\|_{1} (173)
USϱSRUSVSΨSRVS1\displaystyle\leq\|U^{S}\varrho^{SR}U^{S\dagger}-V^{S}\Psi^{SR}V^{S\dagger}\|_{1} (174)
2(USVS)|ϱSR2\displaystyle\leq 2\|(U^{S}-V^{S})|\varrho\rangle^{SR}\|_{2} (175)
2ϱS1/2USVS2.\displaystyle\leq 2\|\varrho^{S}\|_{\infty}^{1/2}\|U^{S}-V^{S}\|_{2}. (176)

Here, the second inequality follows from the monotonicity of the trace distance under any trace-non-increasing map, the third one from that ΨSR\Psi^{SR} is a pure state and the exact diagonalization with the use of the inequality 1x22(1x)\sqrt{1-x^{2}}\leq\sqrt{2(1-x)} for any x[0,1]x\in[0,1], and the last one from the fact that

MX|ϕXY22\displaystyle\|M^{X}|\phi\rangle^{XY}\|_{2}^{2} =Tr[MXϕXYMX]\displaystyle=\operatorname{Tr}[M^{X}\phi^{XY}M^{X\dagger}] (177)
=ϕXMXMX1\displaystyle=\|\phi^{X}M^{X\dagger}M^{X}\|_{1} (178)
ϕXMXMX1\displaystyle\leq\|\phi^{X}\|_{\infty}\|M^{X\dagger}M^{X}\|_{1} (179)
=ϕXMX22,\displaystyle=\|\phi^{X}\|_{\infty}\|M^{X}\|_{2}^{2}, (180)

due to the Hölder’s inequality.

Explicitly writing USU^{S} and VSV^{S} as j𝒥UjS\bigoplus_{j\in\mathcal{J}}U_{j}^{S} and j𝒥VjS\bigoplus_{j\in\mathcal{J}}V_{j}^{S}, respectively, USVS2\|U^{S}-V^{S}\|_{2} is given by j𝒥UjSVjS||22\sum_{j\in\mathcal{J}}\bigl{\|}U_{j}^{S}-V_{j}^{S}\bigr{|}\!\bigr{|}_{2}^{2}, i.e. USVS2=D(US,VS)\|U^{S}-V^{S}\|_{2}=D(U^{S},V^{S}). Hence, for any US𝖴×U^{S}\in{\sf U}_{\times} and VS𝖴×V^{S}\in{\sf U}_{\times}, we obtain

|F(UradS,,UJS)F(VradS,,VJS)|D(US,VS)2ϱS,\frac{|F(U^{S}_{\rm rad},\dots,U^{S}_{J})-F(V^{S}_{\rm rad},\dots,V^{S}_{J})|}{D(U^{S},V^{S})}\\ \leq 2\sqrt{\|\varrho^{S}\|_{\infty}}, (181)

which implies that the Lipschitz constant of FF is bounded from above by 2ϱS2\sqrt{\|\varrho^{S}\|_{\infty}}.

We then apply Theorem 6 and obtain

ProbUS𝖧×[F(UradS,,UJS)212Hmin(S|ER)Γ+δ]\displaystyle{\rm Prob}_{U^{S}\sim{\sf H}_{\times}}\bigl{[}F(U^{S}_{\rm rad},\dots,U^{S}_{J})\geq 2^{-\frac{1}{2}H_{\rm min}(S^{*}|ER)_{\Gamma}}+\delta\bigr{]}
ProbUS𝖧×[F(UradS,,UJS)𝔼[F]+δ]\displaystyle\leq{\rm Prob}_{U^{S}\sim{\sf H}_{\times}}\bigl{[}F(U^{S}_{\rm rad},\dots,U^{S}_{J})\geq\mathbb{E}[F]+\delta\bigr{]} (182)
exp[δ2dmin48ϱS],\displaystyle\leq\exp\bigl{[}-\frac{\delta^{2}d_{\rm min}}{48\|\varrho^{S}\|_{\infty}}\bigr{]}, (183)

where dmin=minj𝒥{dj}d_{\rm min}=\min_{j{\mathcal{J}}}\{d_{j}\}, and Theorem 5, stating that 𝔼[F]212Hmin(S|ER)Γ\mathbb{E}[F]\leq 2^{-\frac{1}{2}H_{\rm min}(S^{*}|ER)_{\Gamma}}, is used to obtain the first inequality. \hfill\blacksquare

Appendix B Proof of Corollary 2

To show Corollary 2, we use the gentle measurement lemma:

Lemma 7 (Gentle measurement lemma [64]).

Let Φ\Phi be in 𝒮()\mathcal{S}(\mathcal{H}) and Λ\Lambda be an Hermitian operator such that 0ΛI0\leq\Lambda\leq I. If they satisfy Tr[ΛΦ]1ϵ\operatorname{Tr}[\Lambda\Phi]\geq 1-\epsilon, where 0ϵ10\leq\epsilon\leq 1, then ΦΦ12ϵ\|\Phi-\Phi^{\prime}\|_{1}\leq 2\sqrt{\epsilon}, where

Φ=ΛΦΛTr[ΛΦ].\Phi^{\prime}=\frac{\sqrt{\Lambda}\Phi\sqrt{\Lambda}}{\operatorname{Tr}[\Lambda\Phi]}. (184)

We also use a simple fact about the conditional min-entropy as given in Lemma 8.

Lemma 8 (Conditional min-entropy after projective measurement).

Let ΠA\Pi^{A} be a projection operator, ΨAB\Psi^{AB} be a quantum state. A post-measured state Ψ~AB:=ΠAΨABΠA/Tr[ΠAΨAB]\tilde{\Psi}^{AB}:=\Pi^{A}\Psi^{AB}\Pi^{A}/\operatorname{Tr}[\Pi^{A}\Psi^{AB}] satisfies

Hmin(A|B)Ψ~Hmin(A|B)Ψ+log[Tr[ΠAΨAB]].H_{\rm min}(A|B)_{\tilde{\Psi}}\geq H_{\rm min}(A|B)_{\Psi}+\log[\operatorname{Tr}[\Pi^{A}\Psi^{AB}]]. (185)
Proof (Lemma 8).

Let σB𝒮(B)\sigma^{B}\in{\mathcal{S}}({\mathcal{H}}^{B}) be the state such that 2Hmin(A|B)ΨIAσBΨAB2^{-H_{\rm min}(A|B)_{\Psi}}I^{A}\otimes\sigma^{B}\geq\Psi^{AB}. Then, we have

2(Hmin(A|B)Ψ+log[Tr[ΠAΨAB]])IAσBΨ~AB,2^{-(H_{\rm min}(A|B)_{\Psi}+\log[\operatorname{Tr}[\Pi^{A}\Psi^{AB}]])}I^{A}\otimes\sigma^{B}\geq\tilde{\Psi}^{AB}, (186)

which implies the desired result. \hfill\blacksquare

Using these lemmas, Corollary 2 can be shown as follows.

Proof (Corollary 2).

We first define the state ϱ~SR\tilde{\varrho}^{SR} by

ϱ~SR:=ΠSϱSRΠSTr[ΠSϱSR]\tilde{\varrho}^{SR}:=\frac{\Pi_{\geq}^{S}\varrho_{SR}\Pi_{\geq}^{S}}{\operatorname{Tr}[\Pi_{\geq}^{S}\varrho_{SR}]} (187)

We also use the fact that ΓER:=TrS[ΓSER]=𝔼US𝖧×[SE(USϱSRUS)]\Gamma^{ER}:=\operatorname{Tr}_{S^{*}}[\Gamma^{S^{*}ER}]=\mathbb{E}_{U^{S}\sim{\sf H}_{\times}}\bigl{[}{\mathcal{M}}^{S\rightarrow E}\bigl{(}U^{S}\varrho^{SR}U^{S\dagger}\bigr{)}\bigr{]}. Using the triangle inequality, we obtain

SE(USϱSRUS)𝔼US𝖧×[SE(USϱSRUS)]1SE(USϱSRUSUSϱ~SRUS)1+SE(USϱ~SRUS)𝔼US𝖧×[SE(USϱ~SRUS)]1+𝔼US𝖧×[SE(USϱ~SRUS)]𝔼US𝖧×[SE(USϱSRUS)]1.\bigl{\|}{\mathcal{M}}^{S\rightarrow E}\bigl{(}U^{S}\varrho^{SR}U^{S\dagger}\bigr{)}-\mathbb{E}_{U^{S}\sim{\sf H}_{\times}}\bigl{[}{\mathcal{M}}^{S\rightarrow E}\bigl{(}U^{S}\varrho^{SR}U^{S\dagger}\bigr{)}\bigr{]}\bigr{\|}_{1}\\ \leq\bigl{\|}{\mathcal{M}}^{S\rightarrow E}\bigl{(}U^{S}\varrho^{SR}U^{S\dagger}-U^{S}\tilde{\varrho}^{SR}U^{S\dagger}\bigr{)}\bigr{\|}_{1}+\bigl{\|}{\mathcal{M}}^{S\rightarrow E}\bigl{(}U^{S}\tilde{\varrho}^{SR}U^{S\dagger}\bigr{)}-\mathbb{E}_{U^{S}\sim{\sf H}_{\times}}\bigl{[}{\mathcal{M}}^{S\rightarrow E}\bigl{(}U^{S}\tilde{\varrho}^{SR}U^{S\dagger}\bigr{)}\bigr{]}\bigr{\|}_{1}\\ +\bigl{\|}\mathbb{E}_{U^{S}\sim{\sf H}_{\times}}\bigl{[}{\mathcal{M}}^{S\rightarrow E}\bigl{(}U^{S}\tilde{\varrho}^{SR}U^{S\dagger}\bigr{)}\bigr{]}-\mathbb{E}_{U^{S}\sim{\sf H}_{\times}}\bigl{[}{\mathcal{M}}^{S\rightarrow E}\bigl{(}U^{S}\varrho^{SR}U^{S\dagger}\bigr{)}\bigr{]}\bigr{\|}_{1}. (188)

In the following, we evaluate each term in the right-hand side of Eq. (188) separately.

For the first term, noting that 𝒞SE{\cal C}^{S\rightarrow E} is a trace-non-increasing map, and the trace norm is unitarily invariant, we have

SE(USϱSRUSUSϱ~SRUS)1\displaystyle\bigl{\|}{\mathcal{M}}^{S\rightarrow E}\bigl{(}U^{S}\varrho^{SR}U^{S\dagger}-U^{S}\tilde{\varrho}^{SR}U^{S\dagger}\bigr{)}\bigr{\|}_{1}
ϱSRϱ~SR1\displaystyle\leq\bigl{\|}\varrho^{SR}-\tilde{\varrho}^{SR}\bigr{\|}_{1} (189)
2ϵ,\displaystyle\leq 2\sqrt{\epsilon}, (190)

where the last inequality follows from the gentle measurement lemma and the assumption that Tr[ϱSRΠS]1ϵ\operatorname{Tr}[\varrho^{SR}\Pi_{\geq}^{S}]\geq 1-\epsilon.

To evaluate the second term, we use Theorem 1. Recalling that Ψ~S\tilde{\Psi}^{S} does not have support on the subspace jS{\mathcal{H}}_{j}^{S} with dimension being smaller than dthd_{\rm th}, it follows that, for any δ>0\delta>0,

SE(USϱ~SRUS)𝔼US𝖧×[SE(USϱ~SRUS)]1212Hmin(S|ER)Γ~+δ\bigl{\|}{\mathcal{M}}^{S\rightarrow E}\bigl{(}U^{S}\tilde{\varrho}^{SR}U^{S\dagger}\bigr{)}-\mathbb{E}_{U^{S}\sim{\sf H}_{\times}}\bigl{[}{\mathcal{M}}^{S\rightarrow E}\bigl{(}U^{S}\tilde{\varrho}^{SR}U^{S\dagger}\bigr{)}\bigr{]}\bigr{\|}_{1}\leq 2^{-\frac{1}{2}H_{\rm min}(S^{*}|ER)_{\tilde{\Gamma}}}+\delta (191)

with probability at least 1exp[δ2dth48ϱ~S]1-\exp[-\frac{\delta^{2}d_{\rm th}}{48\|\tilde{\varrho}^{S}\|_{\infty}}]. Here, Γ~SER=j,j𝒥DSdjdjζjjSEϱ~jjSR\tilde{\Gamma}^{S^{*}ER}=\sum_{j,j^{\prime}\in\mathcal{J}}\frac{D_{S}}{\sqrt{d_{j}d_{j^{\prime}}}}\zeta^{SE}_{jj^{\prime}}\otimes\tilde{\varrho}^{S^{\prime}R}_{jj^{\prime}}. Since Π\Pi_{\geq} is commutable with ΠjS\Pi_{j}^{S} for any jj,

Γ~SER=(ISΠS)ΓSER(ISΠS)Tr[(ISΠS)ΓSER].\tilde{\Gamma}^{S^{*}ER}=\frac{(I^{S}\otimes\Pi^{S^{\prime}}_{\geq})\Gamma^{S^{*}ER}(I^{S}\otimes\Pi^{S^{\prime}}_{\geq})}{\operatorname{Tr}[(I^{S}\otimes\Pi^{S^{\prime}}_{\geq})\Gamma^{S^{*}ER}]}. (192)

Using Lemma 8, we have

Hmin(S|ER)Γ~\displaystyle H_{\rm min}(S^{*}|ER)_{\tilde{\Gamma}}
Hmin(S|ER)Γ+log[Tr[ΠSΓSER]]\displaystyle\geq H_{\rm min}(S^{*}|ER)_{\Gamma}+\log\bigl{[}\operatorname{Tr}[\Pi^{S}_{\geq}\Gamma^{S^{*}ER}]\bigr{]} (193)
=Hmin(S|ER)Γ+log[Tr[ΠSϱSR]]\displaystyle=H_{\rm min}(S^{*}|ER)_{\Gamma}+\log\bigl{[}\operatorname{Tr}[\Pi^{S}_{\geq}\varrho^{SR}]\bigr{]} (194)
Hmin(S|ER)Γ+log[1ϵ],\displaystyle\geq H_{\rm min}(S^{*}|ER)_{\Gamma}+\log[1-\epsilon], (195)

where the second line is obtained since DSdjdjTr[ζjjSE]=δjj\frac{D_{S}}{\sqrt{d_{j}d_{j^{\prime}}}}\operatorname{Tr}[\zeta^{SE}_{jj^{\prime}}]=\delta_{jj^{\prime}}. Furthermore, it holds that

ϱ~S\displaystyle\|\tilde{\varrho}^{S}\|_{\infty} min{1,ΠSϱSTr[ΠSϱS]}\displaystyle\leq\min\bigl{\{}1,\frac{\|\Pi_{\geq}^{S}\|_{\infty}\|\varrho^{S}\|_{\infty}}{\operatorname{Tr}[\Pi^{S}_{\geq}\varrho^{S}]}\bigr{\}} (196)
min{1,ϱS1ϵ}=:C,\displaystyle\leq\min\bigl{\{}1,\frac{\|\varrho^{S}\|_{\infty}}{1-\epsilon}\bigr{\}}=:C, (197)

where we have used the sub-multiplicativity of the operator norm, and the assumption that Tr[ϱSRΠS]1ϵ\operatorname{Tr}[\varrho^{SR}\Pi_{\geq}^{S}]\geq 1-\epsilon. Combining all of these together, the second term is bounded as

SE(USϱ~SRUS)𝔼US𝖧×[SE(USϱ~SRUS)]1212Hmin(S|ER)Γ1ϵ+δ\bigl{\|}{\mathcal{M}}^{S\rightarrow E}\bigl{(}U^{S}\tilde{\varrho}^{SR}U^{S\dagger}\bigr{)}-\mathbb{E}_{U^{S}\sim{\sf H}_{\times}}\bigl{[}{\mathcal{M}}^{S\rightarrow E}\bigl{(}U^{S}\tilde{\varrho}^{SR}U^{S\dagger}\bigr{)}\bigr{]}\bigr{\|}_{1}\leq\frac{2^{-\frac{1}{2}H_{\rm min}(S^{*}|ER)_{\Gamma}}}{\sqrt{1-\epsilon}}+\delta (198)

for any δ>0\delta>0 with probability at least 1exp[δ2dth48C]1-\exp[-\frac{\delta^{2}d_{\rm th}}{48C}].

To evaluate the third term in the right-hand side of Eq. (188), we use the explicit form of the averaged operator

𝔼US𝖧×[SE(USΨSRUS)]=j𝒥DSdjζjjEϱjjR.\mathbb{E}_{U^{S}\sim{\sf H}_{\times}}\bigl{[}{\mathcal{M}}^{S\rightarrow E}\bigl{(}U^{S}\Psi^{SR}U^{S\dagger}\bigr{)}\bigr{]}\\ =\sum_{j\notin\mathcal{J}_{\geq}}\frac{D_{S}}{d_{j}}\zeta^{E}_{jj}\otimes\varrho^{R}_{jj}. (199)

Further, using the relation ΠS=j𝒥ΠjS\Pi^{S}_{\geq}=\sum_{j\in{\mathcal{J}}_{\geq}}\Pi^{S}_{j}, we obtain that the third term XX in the right-hand side of Eq. (188) satisfy

X\displaystyle X =j𝒥DSdjζjjE(ϱ~jjRϱjjR)1\displaystyle=\bigl{\|}\sum_{j\in\mathcal{J}}\frac{D_{S}}{d_{j}}\zeta^{E}_{jj}\otimes(\tilde{\varrho}^{R}_{jj}-\varrho^{R}_{jj})\bigr{\|}_{1} (200)
j𝒥DSdjζjjEϱjjR1\displaystyle\leq\sum_{j\notin\mathcal{J}_{\geq}}\bigl{\|}\frac{D_{S}}{d_{j}}\zeta^{E}_{jj}\otimes\varrho^{R}_{jj}\bigr{\|}_{1}
+j𝒥DSdjζjjE(ϱ~jjRϱjjR)1.\displaystyle\hskip 28.45274pt+\sum_{j\in\mathcal{J}_{\geq}}\bigl{\|}\frac{D_{S}}{d_{j}}\zeta^{E}_{jj}\otimes(\tilde{\varrho}^{R}_{jj}-\varrho^{R}_{jj})\bigr{\|}_{1}. (201)

Recalling that Tr[DSdjζjjE]1\operatorname{Tr}[\frac{D_{S}}{d_{j}}\zeta^{E}_{jj}]\leq 1 and that ϱ~jR=ϱjR/Tr[ΠSϱSR]\tilde{\varrho}^{R}_{j}=\varrho^{R}_{j}/\operatorname{Tr}[\Pi_{\geq}^{S}\varrho^{SR}] for j𝒥j\in{\mathcal{J}}_{\geq}, the first term in the right-hand side of Eq. (201) is bounded from above by

j𝒥Tr[ϱjjR]=Tr[ϱSR(ISΠS)]ϵ.\sum_{j\notin\mathcal{J}_{\geq}}\operatorname{Tr}[\varrho^{R}_{jj}]=\operatorname{Tr}\bigl{[}\varrho^{SR}(I^{S}-\Pi^{S}_{\geq})\bigr{]}\leq\epsilon. (202)

An upper bound of the second term in the right-hand side of Eq. (201) is given by

|1Tr[ΠSϱSR]1|j𝒥Tr[ϱjjSR]\displaystyle\biggl{|}\frac{1}{\operatorname{Tr}[\Pi^{S}_{\geq}\varrho^{SR}]}-1\biggr{|}\sum_{j\in\mathcal{J}_{\geq}}\operatorname{Tr}[\varrho^{SR}_{jj}]
ϵ1ϵj𝒥Tr[ϱjjSR]\displaystyle\leq\frac{\epsilon}{1-\epsilon}\sum_{j\in\mathcal{J}}\operatorname{Tr}[\varrho^{SR}_{jj}] (203)
=ϵ1ϵ.\displaystyle=\frac{\epsilon}{1-\epsilon}. (204)

Combining the upper bounds of all three terms in Eq. (188), we obtain the desired result:

SE(USϱSRUS)𝔼US𝖧×[SE(USϱSRUS)]1212Hmin(S|ER)Γ1ϵ+δ+f(ϵ),\bigl{\|}{\mathcal{M}}^{S\rightarrow E}\bigl{(}U^{S}\varrho^{SR}U^{S\dagger}\bigr{)}\\ -\mathbb{E}_{U^{S}\sim{\sf H}_{\times}}\bigl{[}{\mathcal{M}}^{S\rightarrow E}\bigl{(}U^{S}\varrho^{SR}U^{S\dagger}\bigr{)}\bigr{]}\bigr{\|}_{1}\\ \leq\frac{2^{-\frac{1}{2}H_{\rm min}(S^{*}|ER)_{\Gamma}}}{\sqrt{1-\epsilon}}+\delta+f(\epsilon), (205)

with probability at least 1exp[δ2dth48C]1-\exp[-\frac{\delta^{2}d_{\rm th}}{48C}], where f(ϵ)=2ϵ+ϵ+ϵ1ϵf(\epsilon)=2\sqrt{\epsilon}+\epsilon+\frac{\epsilon}{1-\epsilon}. \hfill\blacksquare

Appendix C Computation of pnp_{n}

We here show Eq. (63) by explicitly computing pnp_{n}, which is defined by

pn:=𝔼US𝖧×[pn(US)],\displaystyle p_{n}:=\mathbb{E}_{U^{S}\sim{\sf H}_{\times}}\bigl{[}p_{n}(U^{S})\bigr{]}, (206)
pn(US):=Tr[ΠnSradUS(ΦARξBin)US],\displaystyle p_{n}(U^{S}):=\operatorname{Tr}\bigl{[}\Pi_{n}^{S_{\rm rad}}U^{S}(\Phi^{AR}\otimes\xi^{B_{\rm in}})U^{S\dagger}\bigr{]}, (207)
=Tr[ΠnSradUS(πAξBin)US].\displaystyle\hskip 34.1433pt=\operatorname{Tr}\bigl{[}\Pi_{n}^{S_{\rm rad}}U^{S}(\pi^{A}\otimes\xi^{B_{\rm in}})U^{S\dagger}\bigr{]}. (208)

We first take the average of US(πAξBin)USU^{S}(\pi^{A}\otimes\xi^{B_{\rm in}})U^{S\dagger} over US𝖧×U^{S}\sim{\sf H}_{\times}, which leads to

𝔼US𝖧×[US(ΠAξBin)US]=m=0N+kTr[ΠmS(ΠAξBin)]πmS,\mathbb{E}_{U^{S}\sim{\sf H}_{\times}}\bigl{[}U^{S}(\Pi^{A}\otimes\xi^{B_{\rm in}})U^{S\dagger}\bigr{]}\\ =\sum_{m=0}^{N+k}\operatorname{Tr}\bigl{[}\Pi_{m}^{S}(\Pi^{A}\otimes\xi^{B_{\rm in}})\bigr{]}\pi_{m}^{S}, (209)

due to the Schur’s lemma. By substituting this, we have

pn=m=0N+kTr[ΠmS(πAξBin)]Tr[πmS(ISinΠnSrad)].p_{n}=\sum_{m=0}^{N+k}\operatorname{Tr}\bigl{[}\Pi_{m}^{S}(\pi^{A}\otimes\xi^{B_{\rm in}})\bigr{]}\operatorname{Tr}\bigl{[}\pi_{m}^{S}(I^{S_{\rm in}}\otimes\Pi^{S_{\rm rad}}_{n})\bigr{]}. (210)

It is then straightforward to compute each term. The first trace is given by

Tr[ΠmS(πAξBin)]\displaystyle\operatorname{Tr}\bigl{[}\Pi_{m}^{S}(\pi^{A}\otimes\xi^{B_{\rm in}})\bigr{]} (211)
=κ=0kTr[(ΠκAΠmκBin)(πAξBin)],\displaystyle=\sum_{\kappa=0}^{k}\operatorname{Tr}\bigl{[}(\Pi_{\kappa}^{A}\otimes\Pi_{m-\kappa}^{B_{\rm in}})(\pi^{A}\otimes\xi^{B_{\rm in}})\bigr{]}, (212)
=12kκ=0k(kκ)χmκ,\displaystyle=\frac{1}{2^{k}}\sum_{\kappa=0}^{k}\binom{k}{\kappa}\chi_{m-\kappa}, (213)

where κmκ=Tr[ΠmκBinξBin]\kappa_{m-\kappa}=\operatorname{Tr}[\Pi_{m-\kappa}^{B_{\rm in}}\xi^{B_{\rm in}}]. The second trace can be computed as

Tr[πmS(ISinΠnSrad)]\displaystyle\operatorname{Tr}\bigl{[}\pi_{m}^{S}(I^{S_{\rm in}}\otimes\Pi^{S_{\rm rad}}_{n})\bigr{]}
=1(N+km)Tr[ΠmS(ISinΠnSrad)]\displaystyle=\frac{1}{\binom{N+k}{m}}\operatorname{Tr}\bigl{[}\Pi_{m}^{S}(I^{S_{\rm in}}\otimes\Pi^{S_{\rm rad}}_{n})\bigr{]} (214)
=1(N+km)Tr[ΠmnSinΠnSrad]\displaystyle=\frac{1}{\binom{N+k}{m}}\operatorname{Tr}\bigl{[}\Pi_{m-n}^{S_{\rm in}}\otimes\Pi^{S_{\rm rad}}_{n}\bigr{]} (215)
=1(N+km)(N+kmn)(n).\displaystyle=\frac{1}{\binom{N+k}{m}}\binom{N+k-\ell}{m-n}\binom{\ell}{n}. (216)

In total, we have Eq. (63).

Appendix D Empirical smoothing of the conditional entropy

We here show Proposition 3. The statement is that, for any δ>0\delta>0, the dynamics KerrSSin\mathcal{L}^{S\rightarrow S_{\rm in}}_{\rm Kerr} leads to

TrSrad[USΨSRUS]ΓSinR1Θξδ(N,k,),\bigl{\|}\operatorname{Tr}_{S_{\rm rad}}\bigl{[}U^{S}\Psi^{SR}U^{S\dagger}\bigr{]}-\Gamma^{S_{\rm in}R}\bigr{\|}_{1}\leq\Theta_{\xi}^{\delta}(N,k,\ell), (217)

with probability at least 1exp[δ2dmin(ϵ)/48]1-\exp[-\delta^{2}d_{\rm min}(\epsilon)/48], where US:=mUmSU^{S}:=\bigoplus_{m}U_{m}^{S} and ΨSR:=ΦARξBin\Psi^{SR}:=\Phi^{AR}\otimes\xi^{B_{\rm in}}. For the definitions of Θξδ(N,k,)\Theta^{\delta}_{\xi}(N,k,\ell), see Eq. (91).

Using the triangle inequality, we have

TrSrad[USΨSRUS]ΓSinR1TrSrad[ΨUSR]TrSrad[ΠϵSradΨUSRΠϵSrad]1+TrSrad[ΠϵSradΨUSRΠϵSrad]Γ~SinR(ϵ)1+Γ~SinR(ϵ)ΓSinR1,\bigl{\|}\operatorname{Tr}_{S_{\rm rad}}\bigl{[}U^{S}\Psi^{SR}U^{S\dagger}\bigr{]}-\Gamma^{S_{\rm in}R}\bigr{\|}_{1}\\ \leq\bigl{\|}\operatorname{Tr}_{S_{\rm rad}}[\Psi_{U}^{SR}]-\operatorname{Tr}_{S_{\rm rad}}[\Pi^{S_{\rm rad}}_{\epsilon}\Psi_{U}^{SR}\Pi^{S_{\rm rad}}_{\epsilon}]\bigr{\|}_{1}\\ +\bigl{\|}\operatorname{Tr}_{S_{\rm rad}}[\Pi^{S_{\rm rad}}_{\epsilon}\Psi_{U}^{SR}\Pi^{S_{\rm rad}}_{\epsilon}]-\tilde{\Gamma}^{S_{\rm in}R}(\epsilon)\bigr{\|}_{1}\\ +\bigl{\|}\tilde{\Gamma}^{S_{\rm in}R}(\epsilon)-\Gamma^{S_{\rm in}R}\bigr{\|}_{1}, (218)

where ΨUSR=USΨSR(US)\Psi_{U}^{SR}=U^{S}\Psi^{SR}(U^{S})^{\dagger}. Using Theorem 1 with the identification of SinS_{\rm in} and EE, the second term is bounded from above as

TrSrad[ΠϵSradΨUSRΠϵSrad]Γ~SinR(ϵ)1212Hmin(S|SinR)Γ~(ϵ)+δ,\bigl{\|}\operatorname{Tr}_{S_{\rm rad}}[\Pi^{S_{\rm rad}}_{\epsilon}\Psi_{U}^{SR}\Pi^{S_{\rm rad}}_{\epsilon}]-\tilde{\Gamma}^{S_{\rm in}R}(\epsilon)\bigr{\|}_{1}\\ \leq 2^{-\frac{1}{2}H_{\rm min}(S^{*}|S_{\rm in}R)_{\tilde{\Gamma}(\epsilon)}}+\delta, (219)

with probability at least 1exp[δ2dmin(ϵ)/48]1-\exp[-\delta^{2}d_{\rm min}(\epsilon)/48] for any δ>0\delta>0. Note that the minimum dimension is given by dmin(ϵ)d_{\rm min}(\epsilon) due to the application of ΠϵSrad\Pi_{\epsilon}^{S_{\rm rad}}.

When the second term of Eq. (218) is small, the first term of Eq. (218) should be also small. To observe this, we use the fact that ΨUSRΠϵSradΨUSRΠϵSrad\Psi_{U}^{SR}\geq\Pi^{S_{\rm rad}}_{\epsilon}\Psi_{U}^{SR}\Pi^{S_{\rm rad}}_{\epsilon}, which leads to

TrSrad[ΨUSR]TrSrad[ΠϵSradΨUSRΠϵSrad]1\displaystyle\bigl{\|}\operatorname{Tr}_{S_{\rm rad}}[\Psi_{U}^{SR}]-\operatorname{Tr}_{S_{\rm rad}}[\Pi^{S_{\rm rad}}_{\epsilon}\Psi_{U}^{SR}\Pi^{S_{\rm rad}}_{\epsilon}]\bigr{\|}_{1} (220)
=|Tr[ΨUSR]Tr[ΠϵSradΨUSRΠϵSrad]|\displaystyle=|\operatorname{Tr}[\Psi_{U}^{SR}]-\operatorname{Tr}[\Pi^{S_{\rm rad}}_{\epsilon}\Psi_{U}^{SR}\Pi^{S_{\rm rad}}_{\epsilon}]| (221)
=1Tr[ΠϵSradΨUSRΠϵSrad].\displaystyle=1-\operatorname{Tr}[\Pi^{S_{\rm rad}}_{\epsilon}\Psi_{U}^{SR}\Pi^{S_{\rm rad}}_{\epsilon}]. (222)

Using the monotonicity of the trace distance, we also obtain that

|Tr[ΠϵSradΨUSRΠϵSrad]Tr[Γ~SinR(ϵ)]|212Hmin(S|SinR)Γ~(ϵ)+δ.\bigl{|}\operatorname{Tr}[\Pi_{\epsilon}^{S_{\rm rad}}\Psi_{U}^{SR}\Pi_{\epsilon}^{S_{\rm rad}}]-\operatorname{Tr}[\tilde{\Gamma}^{S_{\rm in}R}(\epsilon)]\bigr{|}\\ \leq 2^{-\frac{1}{2}H_{\rm min}(S^{*}|S_{\rm in}R)_{\tilde{\Gamma}(\epsilon)}}+\delta. (223)

Since Γ~SinR(ϵ)=TrS[ΓSER(ϵ)]\tilde{\Gamma}^{S_{\rm in}R}(\epsilon)=\operatorname{Tr}_{S^{*}}[\Gamma^{S^{*}ER}(\epsilon)], we have Tr[Γ~SinR(ϵ)]=Tr[Γ~(ϵ)]=1w(ϵ)\operatorname{Tr}[\tilde{\Gamma}^{S_{\rm in}R}(\epsilon)]=\operatorname{Tr}[\tilde{\Gamma}(\epsilon)]=1-w(\epsilon) and so,

Tr[ΠϵSradΨUSRΠϵSrad]1w(ϵ)212Hmin(S|SinR)Γ~(ϵ)δ.\operatorname{Tr}[\Pi_{\epsilon}^{S_{\rm rad}}\Psi_{U}^{SR}\Pi_{\epsilon}^{S_{\rm rad}}]\\ \geq 1-w(\epsilon)-2^{-\frac{1}{2}H_{\rm min}(S^{*}|S_{\rm in}R)_{\tilde{\Gamma}(\epsilon)}}-\delta. (224)

Thus, when Eq. (219) holds, it also holds that

TrSrad[ΨUSR]TrSrad[ΠϵSradΨUSRΠϵSrad]1212Hmin(S|SinR)Γ~(ϵ)+w(ϵ)+δ.\bigl{\|}\operatorname{Tr}_{S_{\rm rad}}[\Psi_{U}^{SR}]-\operatorname{Tr}_{S_{\rm rad}}[\Pi^{S_{\rm rad}}_{\epsilon}\Psi_{U}^{SR}\Pi^{S_{\rm rad}}_{\epsilon}]\bigr{\|}_{1}\\ \leq 2^{-\frac{1}{2}H_{\rm min}(S^{*}|S_{\rm in}R)_{\tilde{\Gamma}(\epsilon)}}+w(\epsilon)+\delta. (225)

Note however that this evaluation is not tight since, when ϵ=0\epsilon=0, the L.H.S. is trivially zero, but the R.H.S. is in general non-zero.

For the third term, note that ΓSSinRΓ~(ϵ)SSinR\Gamma^{S^{*}S_{\rm in}R}\geq\tilde{\Gamma}(\epsilon)^{S^{*}S_{\rm in}R}, implying that ΓSinRΓ~SinR(ϵ)\Gamma^{S_{\rm in}R}\geq\tilde{\Gamma}^{S_{\rm in}R}(\epsilon). Thus, we have

Γ~SinR(ϵ)ΓSinR1\displaystyle\|\tilde{\Gamma}^{S_{\rm in}R}(\epsilon)-\Gamma^{S_{\rm in}R}\|_{1} =Tr[ΓSinRΓ~SinR(ϵ)]\displaystyle=\operatorname{Tr}[\Gamma^{S_{\rm in}R}-\tilde{\Gamma}^{S_{\rm in}R}(\epsilon)] (226)
=1Tr[Γ~(ϵ)]\displaystyle=1-\operatorname{Tr}[\tilde{\Gamma}(\epsilon)] (227)
=w(ϵ).\displaystyle=w(\epsilon). (228)

Altogether, we obtain that, for any δ>0\delta>0,

TrSrad[USΨSRUS]ΓSinR12112Hmin(S|SinR)Γ~(ϵ)+w(ϵ)+2δ\bigl{\|}\operatorname{Tr}_{S_{\rm rad}}\bigl{[}U^{S}\Psi^{SR}U^{S\dagger}\bigr{]}-\Gamma^{S_{\rm in}R}\bigr{\|}_{1}\\ \leq 2^{1-\frac{1}{2}H_{\rm min}(S^{*}|S_{\rm in}R)_{\tilde{\Gamma}(\epsilon)}}+w(\epsilon)+2\delta (229)

with probability at least 1exp[δ2dmin/48]1-\exp[-\delta^{2}d_{\rm min}/48]. Since the R.H.S. is equal to Θξδ(N,k,)\Theta_{\xi}^{\delta}(N,k,\ell), this completes the proof.

Appendix E Derivation of ηξ\eta_{\xi}

Here, we derive Eq. (86), i.e.,

ηξ(N,k,)\displaystyle\eta_{\xi}(N,k,\ell)
:=ΓSinRΓSinπR1\displaystyle:=\bigl{\|}\Gamma^{S_{\rm in}R}-\Gamma^{S_{\rm in}}\otimes\pi^{R}\bigr{\|}_{1} (230)
=12kν=0N+kκ=0kFκ,ν(ξ)(N+kν)(kκ),\displaystyle=\frac{1}{2^{k}}\sum_{\nu=0}^{N+k-\ell}\sum_{\kappa=0}^{k}F_{\kappa,\nu}(\xi)\binom{N+k-\ell}{\nu}\binom{k}{\kappa}, (231)

where

Fκ,ν(ξ)=|m=0N+k(mν)(N+km)(χmκ12kκ=0k(kκ)χmκ)|,F_{\kappa,\nu}(\xi)\\ =\biggl{|}\sum_{m=0}^{N+k}\frac{\binom{\ell}{m-\nu}}{\binom{N+k}{m}}\biggl{(}\chi_{m-\kappa}-\frac{1}{2^{k}}\sum_{\kappa^{\prime}=0}^{k}\binom{k}{\kappa^{\prime}}\chi_{m-\kappa^{\prime}}\biggr{)}\biggr{|}, (232)

and χμ=Tr[ξBinΠμBin]\chi_{\mu}=\operatorname{Tr}[\xi^{B_{\rm in}}\Pi^{B_{\rm in}}_{\mu}].

It is straightforward to compute ΓSinR\Gamma^{S_{\rm in}R} and ΓSin\Gamma^{S_{\rm in}} as

ΓSinR=\displaystyle\Gamma^{S_{\rm in}R}=
m=0N+kn=0κ=0k(n)(kκ)(N+kmn)(N+km)χmκπnmSinπκR,\displaystyle\ \ \sum_{m=0}^{N+k}\sum_{n=0}^{\ell}\sum_{\kappa=0}^{k}\frac{\binom{\ell}{n}\binom{k}{\kappa}\binom{N+k-\ell}{m-n}}{\binom{N+k}{m}}\chi_{m-\kappa}\pi^{S_{\rm in}}_{n-m}\otimes\pi_{\kappa}^{R}, (233)
ΓSin=12km=0N+kn=0κ=0k(n)(kκ)(N+kmn)(N+km)χmκπmnSin,\displaystyle\Gamma^{S_{\rm in}}=\frac{1}{2^{k}}\sum_{m=0}^{N+k}\sum_{n=0}^{\ell}\sum_{\kappa=0}^{k}\frac{\binom{\ell}{n}\binom{k}{\kappa}\binom{N+k-\ell}{m-n}}{\binom{N+k}{m}}\chi_{m-\kappa}\pi_{m-n}^{S_{\rm in}}, (234)

where we have used ΠmS=n=0ΠmnSinΠnSrad\Pi_{m}^{S^{\prime}}=\sum_{n=0}^{\ell}\Pi_{m-n}^{S_{\rm in}^{\prime}}\otimes\Pi_{n}^{S_{\rm rad}^{\prime}} and ΠmS=κ=0kΠκAΠmκR\Pi_{m}^{S}=\sum_{\kappa=0}^{k}\Pi_{\kappa}^{A}\otimes\Pi_{m-\kappa}^{R}. Since both are already diagonal, it is easy to compute their distance and obtain Eq. (231).

Appendix F Derivation of Eq. (93)

We here derive Eq. (93). To restate it, let {jA}\{{\mathcal{H}}^{A}_{j}\} be mutually orthogonal subspaces of A{\mathcal{H}}^{A}, and πjA\pi_{j}^{A} be the completely mixed state on jA{\mathcal{H}}^{A}_{j}. For any state ΛABC\Lambda^{ABC} in the form of j=0JpjπjAρjBC\sum_{j=0}^{J}p_{j}\pi_{j}^{A}\otimes\rho_{j}^{BC}, where ρjBC𝒮(BC)\rho_{j}^{BC}\in\mathcal{S}(\mathcal{H}^{BC}) and {pj}\{p_{j}\} is a probability distribution, it holds that

2Hmin(AB|C)Λj=0Jpjdj2Hmin(B|C)ρj,2^{-H_{\rm min}(AB|C)_{\Lambda}}\leq\sum_{j=0}^{J}\frac{p_{j}}{d_{j}}2^{-H_{\rm min}(B|C)_{\rho_{j}}}, (235)

where dj=dimjAd_{j}=\dim{\mathcal{H}}^{A}_{j}.

By definition of the conditional min-entropy, j[0,J]\forall j\in[0,J], σjC𝒮(C)\exists\sigma_{j}^{C}\in\mathcal{S}(\mathcal{H}^{C}) such that 2Hmin(B|C)ρjIBσjCρjBC2^{-H_{\rm min}(B|C)_{\rho_{j}}}I^{B}\otimes\sigma_{j}^{C}\geq\rho_{j}^{BC}. Hence, we have

IAB(2Hmin(B|C)ρjσjC)ΠjAρjBC.I^{AB}\otimes(2^{-H_{\rm min}(B|C)_{\rho_{j}}}\sigma_{j}^{C})\geq\Pi_{j}^{A}\otimes\rho_{j}^{BC}. (236)

for all j[1,J]j\in[1,J]. This further implies that

Tr[σ~C]IABσ~CTr[σ~C]ΛABC,\operatorname{Tr}[\tilde{\sigma}^{C}]I^{AB}\otimes\frac{\tilde{\sigma}^{C}}{\operatorname{Tr}[\tilde{\sigma}^{C}]}\geq\Lambda^{ABC}, (237)

where σ~C:=j=0Jpj/dj2Hmin(B|C)ρjσjC\tilde{\sigma}^{C}:=\sum_{j=0}^{J}p_{j}/d_{j}2^{-H_{\rm min}(B|C)_{\rho_{j}}}\sigma_{j}^{C} is an un-normalized state. This concludes the proof.

Appendix G Comparison with the analysis based on the partial decoupling

Refer to caption
Figure 6: Figures of the delay δΔ(L,δL)\delta\ell_{\Delta}(L,\delta L) for the pure Kerr BH for different ZZ-axis AM LL and different fluctuations δL\delta L. The size kk of the quantum information source AA is set to 11. For simplicity, we set Δ=0.1\Delta=0.1. In each plot, the solid line shows the fitting by the function ax+bx+cax+b\sqrt{x}+c, where (a,b,c)(a,b,c) is decided by the least squared fitting.
Refer to caption
Figure 7: Figures for the delay δΔ(L,δL)\delta\ell_{\Delta}(L,\delta L) for the mixed Kerr BH with different ZZ-axis AM LL and fluctuations δL\delta L of the initial Kerr BH BinB_{\rm in}. They are plotted as a function of the number NN of qubits in the initial Kerr BH. The size kk of the quantum information source AA is set to 11. The Δ\Delta is set to 0.10.1 for simplicity.

Here, we provide numerical evaluations of the scaling of the delay δΔ(L,δL)\delta\ell_{\Delta}(L,\delta L) with respect to the size NN of the initial Kerr BH BinB_{\rm in} for various ZZ-axis AMs LL and their fluctuations δL\delta L. The cases of pure and mixed Kerr BHs are depicted in Figs. 6 and 7, respectively.

For the pure Kerr BH, we first observe from the case of δL=0\delta L=0 (black circles) that δΔ(L,0)=O(N)\delta\ell_{\Delta}(L,0)=O(\sqrt{N}) for any LL. Note that the delay is in comparison with the case of L=δL=0L=\delta L=0. Hence, the delay for δL=0\delta L=0 is not plotted in Fig. 6 (I). When δL0\delta L\neq 0, the scaling seems to depend on if δL=O(N)\delta L=O(\sqrt{N}) or O(N)O(N). For δ=O(N)\delta=O(\sqrt{N}), it appears that δΔ(L,δL)=O(N)\delta\ell_{\Delta}(L,\delta L)=O(\sqrt{N}), while δΔ(L,δL)=O(N)\delta\ell_{\Delta}(L,\delta L)=O(N) for δL=O(N)\delta L=O(N). Thus, we conclude that, for any LL, the delay δΔ(L,δL)\delta\ell_{\Delta}(L,\delta L) gradually changes from O(N)O(\sqrt{N}) to O(N)O(N) as δL\delta L increases from O(N)O(\sqrt{N}) to O(N)O(N).

For the mixed BH, the delay δΔ(L,0)\delta\ell_{\Delta}(L,0) is independent of NN. This seems to be also the case for δL=O(N)\delta L=O(\sqrt{N}). When δL=O(N)\delta L=O(N), δΔ(L,δL)\delta\ell_{\Delta}(L,\delta L) appears to increase as NN increases. It also seems that the scaling of δΔ(L,δL)\delta\ell_{\Delta}(L,\delta L) with respect to NN depends on LL: when L=0L=0, the delay may grow very slowly as NN becomes large, while it seems to scale linearly with NN when |L|N/4|L|\geq N/4.