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

Linear Response Theory of Entanglement Entropy

Yuan-Sheng Wang Department of Physics, Southern University of Science and Technology, Shenzhen 518055, China School of Physical Sciences, University of Science and Technology of China, Hefei 230026, China    Teng Ma Beijing Academy of Quantum Information Sciences, Beijing 100193, China    Man-Hong Yung [email protected] Institute for Quantum Science and Engineering, and Department of Physics, Southern University of Science and Technology, Shenzhen 518055, China
Abstract

Linear response theory (LRT) is a key tool in investigating the quantum matter, for quantum systems perturbed by a weak probe, it connects the dynamics of experimental observable with the correlation function of unprobed equilibrium states. Entanglement entropy(EE) is a measure of quantum entanglement, it is a very important quantity of quantum physics and quantum information science. While EE is not an observable, developing the LRT of it is an interesting thing. In this work, we develop the LRT of von Neumann entropy for an open quantum system. Moreover, we found that the linear response of von Neumann entanglement entropy is determined by the linear response of an observable. Using this observable, we define the Kubo formula and susceptibility of EE, which have the same properties of its conventional counterpart. Through using the LRT of EE, we further found that the linear response of EE will be zero for maximally entangled or separable states, this is a unique feature of entanglement dynamics. A numerical verification of our analytical derivation is also given using XX spin chain model. The LRT of EE provides a useful tool in investigating and understanding EE.

I Introduction

Entanglement provides fundamental insight into quantum physicsBell (1964); Bennett et al. (1993) and quantum information scienceBennett and DiVincenzo (2000). For example, in condensed matter physics, entanglement provides a unique tool for characterizing quantum mechanical many-body phases and new kinds of quantum orderLevin and Wen (2006); Kitaev and Preskill (2006). As a bridge linking quantum statistical mechanics with quantum information, entanglement has been exploited to establish efficient methods for examining many-body interacting systemsOsborne and Nielsen (2002). Deep understanding of the entanglement decoherence is also expected to insights into the quantum measurement and the quantum-classical transitionGisin (1996); Dodd and Halliwell (2004); Dodd (2004). Besides the fundamental physical meaning of entanglement, it is also a resource required for many quantum technologies such as quantum communication, quantum sensingDegen et al. (2017), quantum metrologyGiovannetti et al. (2006, 2011); Chin et al. (2012), etc. The ability to generate, manipulate and detect entanglementZeilinger (1998); Tittel et al. (1998); Deutsch and Ekert (1998); DiVincenzo and Terhal (1998) are bases of the whole field of quantum computing.

The interest in the properties of entanglement has also been extended to understanding its dynamical behaviour, which is a powerful perspective in understanding non-equilibrium quantum dynamicsBellomo et al. (2007); Nahum et al. (2018); Zhu et al. (2019). For instance, the dynamics of entanglement has recently been realized as a useful probe in studying ergodicity and its breakdown in quantum many-body systemsŽnidarič et al. (2008); Bardarson et al. (2012); Kim and Huse (2013); Ho and Abanin (2017). Although great progress has been made in this field, to our knowledge, however, there is still a lack of a general tool in investigating how entanglement entropy (EE) response to an external perturbation.

Conventional linear response theory (LRT) is a general and powerful tool in studying quantum mattersFu and Dudley (1993); Beenakker (1991); Wylie and Sipe (1984); Baroni et al. (1987); Hänggi and Thomas (1982); Savrasov and Savrasov (1996), because it connects the dynamical response of a quantum system to an external probe with the correlation functions of the unperturbed equilibrium state. While the conventional LRT is applied to observable, which is associated with a Hermitian operator, as EE is not an observable, finding the LRT for it seems to be not only interesting but also useful.

In this paper, we develop an LRT of von Neumann entropy for an open quantum system which is subject to external perturbation. We found that the LRT of von Neumann entropy to external perturbation is given by the linear response of an observable, using this observable we define the Kubo formula and susceptibility of EE which have the same properties of its conventional counterpart. Using the LRT of EE, we demonstrate that there is no linear response of von Neumann entropy for non-degenerate systems which are initially at separable states and maximally entangled states. To verify our analytical result, we numerically solve the dynamics of von Neumann entropy of a subsystem in a XX spin chain model. The LRT of EE provides a useful tool in understanding and investigating the dynamics of entanglement.

The paper is organized as follows. In Sec.II, we briefly review the conventional LRT. In Sec.III we derive the LRT of von Neumann entropy, define the Kubo formula and the susceptibility of EE, discuss properties and applications of the Kubo formula and the susceptibility of EE. In Sec.IV we give a numerical verification of our analytical results using XX spin chain model. The concluding Sec. V summarizes the findings. Technical details of some calculations are presented in the appendix.

II Review of conventional linear response theory

II.1 The Kubo formula

Consider a quantum system which is weakly coupled to an external perturbation F(t)H^1F(t)\hat{H}_{1} at t=t0t=t_{0} with time dependence F(t)F(t). The Hamiltonian for such system is given by

H^(t)=H^0+F(t)H^1\displaystyle\hat{H}(t)=\hat{H}_{0}+F(t)\hat{H}_{1} (1)

where H^0\hat{H}_{0} is the Hamiltonian of unperturbed system. In interaction picture (we denote an operator or density matrix in interaction picture through adding ”I” to its subscript), the average of any operator O^I(t)\hat{O}_{I}(t) over state ρI(t)\rho_{I}(t) reads

O^I(t)\displaystyle\langle\hat{O}_{I}(t)\rangle tr[ρI(t)O^I(t)]\displaystyle\equiv\text{tr}[\rho_{I}(t)\hat{O}_{I}(t)] (2)

where ρI(t)=U^I(t,t0)ρI(t0)U^I(t,t0)\rho_{I}(t)=\hat{U}_{I}(t,t_{0})\rho_{I}(t_{0})\hat{U}_{I}^{\dagger}(t,t_{0}) with the time evolution operator U^I(t,t0)=𝒯exp[it0t𝑑tF(t)H^1,I(t)]\hat{U}_{I}(t,t_{0})=\mathcal{T}\exp[-i\int_{t_{0}}^{t}dt^{\prime}F(t^{\prime})\hat{H}_{1,I}(t^{\prime})]. If F(t)F(t) is so small such that F(t)H^1F(t)\hat{H}_{1} is much smaller than H^0\hat{H}_{0}, we can describe O^I(t)tr[ρI(t)O^I(t)]=tr[ρ(t)O^(t)]\langle\hat{O}_{I}(t)\rangle\equiv\text{tr}[\rho_{I}(t)\hat{O}_{I}(t)]=\text{tr}[\rho(t)\hat{O}(t)] by performing an expansion in powers of F(t)F(t), working to the first order solution to U^I(t)\hat{U}_{I}(t) (we assume =1\hbar=1 hereafter):

δO^I(t)=t0dtR(t,t)F(t))+𝒪(F2)\displaystyle\delta\langle\hat{O}_{I}(t)\rangle=\int_{t_{0}}^{\infty}dt^{\prime}R(t,t^{\prime})F(t^{\prime}))+\mathcal{O}(F^{2}) (3)

where δO^I(t)O^I(t)FO^I(t)F=0\delta\langle\hat{O}_{I}(t)\rangle\equiv\langle\hat{O}_{I}(t)\rangle_{F}-\langle\hat{O}_{I}(t)\rangle_{F=0} and R(t,t)=iθ(tt))[O^I(t),H^1,I(t)]0R(t,t^{\prime})=-i\theta(t-t^{\prime}))\langle[\hat{O}_{I}(t),\hat{H}_{1,I}(t^{\prime})]\rangle_{0}, with 0tr[ρ(t0)]\langle\cdot\rangle_{0}\equiv\text{tr}\left[\cdot\rho(t_{0})\right] an average with respect to the initial state ρ(t0)\rho(t_{0}). The real function R(t,t)R(t,t^{\prime}) is the well-known Kubo formula of observable O^\hat{O} for system which is subject to a perturbation F(t)H^1F(t)\hat{H}_{1}, which is also referred to as the linear response function. It is readily to see that the Kubo formula can be written in terms of two-point correlation function C(t,t)O^I(t)H^1,I(t)0C(t,t^{\prime})\equiv\langle\hat{O}_{I}(t)\hat{H}_{1,I}(t^{\prime})\rangle_{0}, if ρ(t0)\rho(t_{0}) is an equilibrium state, the correlation function provides a measure of equilibrium fluctuations in the system, therefore, it is revealed by Eq.(3) that although the response of an equilibrium system to external perturbation should move the system away from equilibrium, if the perturbation is weak enough, the response is dictated by the equilibrium fluctuations. We may conclude that the knowledge of the equilibrium state is useful in predicting the behaviour of nonequilibrium process.

II.2 Some properties and applications of the Kubo formula

Stationary. If the initial state is equilibrium state, its density matrix then commutes with H^0\hat{H}_{0}, inserting O^I(t)=eiH^0tO^eiH^0t\hat{O}_{I}(t)=e^{i\hat{H}_{0}t}\hat{O}e^{-i\hat{H}_{0}t} and H^1,I(t)=eiH^0tH^1eiH^0t\hat{H}_{1,I}(t)=e^{i\hat{H}_{0}t^{\prime}}\hat{H}_{1}e^{-i\hat{H}_{0}t^{\prime}} into the expression of correlation function and using the cyclic property of trace, we have

C(t,t)\displaystyle C(t,t^{\prime}) =O^I(tt)H^1\displaystyle=\langle\hat{O}_{I}(t-t^{\prime})\hat{H}_{1}\rangle (4)

that is, the equilibrium state correlation function is invariant under time translations: C(t,t)=C(tt)C(t,t^{\prime})=C(t-t^{\prime}), which results in translation invariant Kubo formula: R(t,t)=R(tt)R(t,t^{\prime})=R(t-t^{\prime}).

Causal. The system should not respond before the external perturbation is applied, therefore R(t)=0R(t)=0 for t<0t<0, this is enforced by the step function in the expression of R(t)R(t). The causality will result in analyticity of χ(ω)\chi(\omega) in the upper half complex plane, we will show it in the following. The Fourier expansion of R(t)R(t) reads

R(t)=12πχ(ω)eiωt𝑑ω\displaystyle R(t)=\frac{1}{2\pi}\int_{-\infty}^{\infty}\chi(\omega)e^{-i\omega t}d\omega (5)

if t<0t<0 we can perform the integral by completing the contour in the upper half plane, the complementary integral has to be zero since the real part of the exponent i(i|ωi|)(|t|)-i(i|\omega_{i}|)\cdot(-|t|)\rightarrow-\infty along the contour in the upper half plane, with ωi\omega_{i} the imaginary part of ω\omega. On the other hand, according to Cauchy’s residue theorem, the integral along the contour is given by the sum of the residues inside the contour. So there is no poles of χ(ω)\chi(\omega) for Imχ(ω)>0\text{Im}\chi(\omega)>0, which means that χ(ω)\chi(\omega) is analytic in the upper half plane.  
Kramers-Krönig relation. Because χ(ω)\chi(\omega) is analytic in the upper half complex plane, it is found that the imaginary and real parts of χ(ω)\chi(\omega) is not independent, they are connected by the Kramers-Krönig relationToll (1956); de L. Kronig (1926); Kramers (1927):

Re[χ(ω)]\displaystyle\text{Re}[\chi(\omega)] =1π𝒫Im[χ(ω)]ωω𝑑ω\displaystyle=\frac{1}{\pi}\mathcal{P}\int_{-\infty}^{\infty}\frac{\text{Im}[\chi(\omega)]}{\omega^{\prime}-\omega}d\omega^{\prime} (6)
Im[χ(ω)]\displaystyle\text{Im}[\chi(\omega)] =1π𝒫Re[χ(ω)]ωω𝑑ω.\displaystyle=-\frac{1}{\pi}\mathcal{P}\int_{-\infty}^{\infty}\frac{\text{Re}[\chi(\omega)]}{\omega^{\prime}-\omega}d\omega^{\prime}. (7)

where 𝒫\mathcal{P} denotes the Cauchy principal value. Using Kramers-Krönig relation, the response function can be reconstructed given just real or imaginary part.

Susceptibility. According to the stationary property, for equilibrium initial state, the Kubo formula is time-translation invariant: R(t,t)=R(tt)R(t,t^{\prime})=R(t-t^{\prime}), Eq.(3) becomes a convolution, its Fourier transform reads

δO^I(ω)=χ(ω)F(ω)+𝒪(F2)\displaystyle\langle\delta\hat{O}_{I}(\omega)\rangle=\chi(\omega)F(\omega)+\mathcal{O}(F^{2}) (8)

where F(ω)F(\omega) is the Fourier transforms of F(t)F(t), χ(ω)\chi(\omega) is the Fourier transform of Kubo formula and is oftern referred to as generalized susceptibility. We learn from Eq.(8) that the response of observable O^\hat{O} to time-dependent perturbation F(t)H^1F(t)\hat{H}_{1} is ’local’ in frequency space, this is a feature of equilibrium linear response. Thanks to this locality, we can directly infer the generalized susceptibility χ(ω)\chi(\omega) of O^\hat{O} for a known perturbation F(t)H^1F(t)\hat{H}_{1} through measuring the time evolution of O^I(t)\langle\hat{O}_{I}(t)\rangle. The susceptibility of some observable to certain perturbation is essential in investigating the properties of equilibrium systems (e.g. classifying different materials using magnetic susceptibility). Therefore the conventional LRT is an important tool in investigating the properties of unknown systems.

Fluctuation-dissipation relation. Denoting the imaginary part of generalized susceptibility χ(ω)\chi(\omega) as χ′′(ω)\chi^{\prime\prime}(\omega), it can be written as

χ′′(ω)=12i𝑑teiωt[R(t)R(t)]\displaystyle\chi^{\prime\prime}(\omega)=\frac{1}{2i}\int_{-\infty}^{\infty}dte^{i\omega t}[R(t)-R(-t)] (9)

which is not invariant under time-reversal ttt\rightarrow-t. Hence, χ′′(ω)\chi^{\prime\prime}(\omega) is called the dissipative or absorptive part of the generalized susceptibility. The fluctuation–dissipation theorem says that the dissipation effects are determined by natural fluctuation in thermal equilibrium state. To be more specific, for canonical ensemble with the inverse temperature β\beta, ρ=eβH^0/Z\rho=e^{-\beta\hat{H}_{0}}/Z, the imaginary part of susceptibility χ′′(ω)\chi^{\prime\prime}(\omega) which describe the dissipation and (the Fourier transform of) the correlation function C~(ω)\tilde{C}(\omega) are connected by the Fluctuation-dissipation relation:

C~(ω)=2[nB(ω)+1]χ′′(ω)\displaystyle\tilde{C}(\omega)=-2[n_{B}(\omega)+1]\chi^{\prime\prime}(\omega) (10)

where nB(ω)=1/(eβω1)n_{B}(\omega)=1/(e^{\beta\omega}-1) is the Bose-Einstein distribution function.

III The Linear Response Theory of Entanglement Entropy

III.1 The Kubo formula of EE

Be similar to the conventional case, supposing at time t=t0t=t_{0}, a time-dependent perturbation F(t)H^1F(t)\hat{H}_{1} is applied to a composite system which consists of subsystems A and B, its Hamiltonian has the same form as (1). Solving the dynamics of this composite system using perturbative method (see appendix A for detail), we have

δρ(t)=it0t𝑑τ[H^1,I(τt),ρ0]F(τ)+𝒪(F2),\displaystyle\delta\rho(t)=-i\int_{t_{0}}^{t}d\tau[\hat{H}_{1,I}(\tau-t),\rho_{0}]F(\tau)+\mathcal{O}(F^{2}), (11)

where ρ0ρ(t0)\rho_{0}\equiv\rho(t_{0}), δρ(t)ρ(t)ρ0\delta\rho(t)\equiv\rho(t)-\rho_{0}, and H^1,I(t)=U^0(t,t0)H^1U^0(t,t0)\hat{H}_{1,I}(t)=\hat{U}_{0}^{\dagger}(t,t_{0})\hat{H}_{\text{1}}\hat{U}_{0}(t,t_{0}), with U^0(t,t0)=exp[iH^0(tt0)]\hat{U}_{0}(t,t_{0})=\exp[-i\hat{H}_{0}(t-t_{0})].

The entanglement entropy of subsystem A is given by SA(t)=tr[ρA(t)lnρA(t)]S_{A}(t)=-\text{tr}[\rho_{A}(t)\ln\rho_{A}(t)], with ρA(t)=trB[ρ(t)]\rho_{A}(t)=\text{tr}_{B}[\rho(t)]. For infinitesimal change δρA(t)ρA(t)ρA(t0)\delta\rho_{A}(t)\equiv\rho_{A}(t)-\rho_{A}(t_{0}), the time-dependent change in entanglement entropy readsMa et al. (2019): δSA(t)SA(t)SA(t0)trA[δρA(t)lnρA(t)]\delta S_{A}(t)\equiv S_{A}(t)-S_{A}(t_{0})\approx-\text{tr}_{A}\left[\delta\rho_{A}(t)\ln\rho_{A}(t)\right], furthermore, we can expand the operator lnρA(t)\ln\rho_{A}(t) in terms of δρA(t)\delta\rho_{A}(t) as lnρA(t)=lnρA(t0)|+𝒪(δρA)\ln\rho_{A}(t)=\ln\rho_{A}(t_{0})|+\mathcal{O}(\delta\rho_{A}), then to the first order in δρA(t)\delta\rho_{A}(t), we have

δSA(t)=trA[δρA(t)lnρA(t0)]+𝒪(δρA).\displaystyle\delta S_{A}(t)=-\text{tr}_{A}\left[\delta\rho_{A}(t)\ln\rho_{A}(t_{0})\right]+\mathcal{O}(\delta\rho_{A}). (12)

What’s interesting about this equation is that the right-hand side of (12) has the form of the change in the expectation value of an Hermitian operator lnρA(t0)\ln\rho_{A}(t_{0}). Similar result can be seen in Ref. Strasberg et al. (2017).

Inserting Eq. (11) into Eq. (12), we obtain the time evolution of EE under weak driving (see appendix B for detail):

δSA(t)=t0RE(t,τ)F(τ)𝑑τ+𝒪(F2)\displaystyle\delta S_{A}(t)=\int_{t_{0}}^{\infty}R_{E}(t,\tau)F(\tau)d\tau+\mathcal{O}(F^{2}) (13)

where RE(t,τ)R_{E}(t,\tau) is given by the following expression

RE(t,τ)=iθ(tτ)[s^A,H^1,I(τt)]0.\displaystyle R_{E}(t,\tau)=-i\theta(t-\tau)\langle[\hat{s}_{A},\hat{H}_{1,I}(\tau-t)]\rangle_{0}. (14)

where s^AlnρA(t0)\hat{s}_{A}\equiv-\ln\rho_{A}(t_{0}), Comparing Eq. (13) and Eq. (14) with Eq. (3), we found that to first order in F(t)F(t), the change in EE is equivalent to the change in the expectation of observable s^A(t)\langle\hat{s}_{A}(t)\rangle:

δSA(t)δs^A(t)\displaystyle\delta S_{A}(t)\approx\delta\langle\hat{s}_{A}(t)\rangle (15)

Eq. (13), Eq. (14) and Eq. (15) are key results of this work, they revealed that: for equilibrium open system A, the response of EE is determined by the linear response of observable s^A=lnρA(t0)\hat{s}_{A}=-\ln\rho_{A}(t_{0}). Eq. (14) has the same form as conventional linear response function, it describes how EE and the observable s^A\hat{s}_{A} response to weak external force F(t)H^1F(t)\hat{H}_{1}, we refer to RE(t,τ)R_{E}(t,\tau) as the Kubo formula of EE.

III.2 Properties and applications of the EE Kubo formula

The Kubo formula of EE is also the Kubo formula of an observable, this fact implies that RE(t,τ)R_{E}(t,\tau) has the same properties of conventional linear response function.

Like conventional LRT, if [ρ0,H^0]=0[\rho_{0},\hat{H}_{0}]=0, the EE Kubo formula will be stationary: RE(t,τ)=RE(tτ)R_{E}(t,\tau)=R_{E}(t-\tau), then the Fourier transform(FT) of Eq.(13) reads

δSA(ω)=χE(ω)F(ω)+𝒪(F2),\displaystyle\delta S_{A}(\omega)=\chi_{E}(\omega)F(\omega)+\mathcal{O}(F^{2}), (16)

where χE(ω)\chi_{E}(\omega) and F(ω)F(\omega) are FT of RE(t)R_{E}(t) and F(t)F(t) respectively. Eq. (16) has the same form as (8) except for that SA(t)S_{A}(t) is not an expectation value of an observable. RE(t)R_{E}(t) and χE(ω)\chi_{E}(\omega) also satisfy properties such as causality, analyticity of χE(ω)\chi_{E}(\omega) in the upper-half plane and there is a Kramers-Kröning relation between real and imaginary part of χE(ω)\chi_{E}(\omega).

Entanglement susceptibility. We refer to χE(ω)\chi_{E}(\omega) as entanglement susceptibility. According to Eq. (13), the time evolution of EE can be determined by measuring an observable s^A(t)\hat{s}_{A}(t). F(t)F(t) is normally known to the one who applies the perturbation, we thus can infer the Kubo formula RE(t)R_{E}(t) from (16) and hence the entanglement susceptibility χE(ω)\chi_{E}(\omega). The Kubo formula is independent of F(t)F(t), for a given H^1(t)\hat{H}_{1}(t), once we determine its Kubo formula, we may predict the response for any F(t)F(t). Although s^A\hat{s}_{A} is hard to measure, for a system with few degrees of freedom, (e. g., a spin-1/2) we can evaluate it using state tomography.

Canonical ensemble EE linear response If the initial state of subsystem A is described by canonical ensemble ρA(t0)=eβH^A/ZA\rho_{A}(t_{0})=e^{-\beta\hat{H}_{A}}/Z_{A} with inverse temperature β=1/kBT\beta=1/k_{B}T, H^A\hat{H}_{A} is the Hamiltonian of subsystem A and ZAZ_{A} is the corresponding partition function, then lnρA(t0)=βH^AlnZA\ln\rho_{A}(t_{0})=-\beta\hat{H}_{A}-\ln Z_{A}, inserting this equation into (13), we have

kBTδSA(t)=t0(1/β)RE(t,τ)F(τ)𝑑τ+𝒪(F2)\displaystyle-k_{B}T\delta S_{A}(t)=\int_{t_{0}}^{\infty}(1/\beta)R_{E}(t,\tau)F(\tau)d\tau+\mathcal{O}(F^{2}) (17)

it is readily to see that (1/β)RE(t,τ)=iθ(tτ)[H^A,H^1,I(τt)]0(1/\beta)R_{E}(t,\tau)=-i\theta(t-\tau)\langle[\hat{H}_{A},\hat{H}_{1,I}(\tau-t)]\rangle_{0}, which is the Kubo formula of observable H^A\hat{H}_{A} and perturbation H^1(t)\hat{H}_{1}(t), δQA(t)kBTδSA\delta Q_{A}(t)\equiv-k_{B}T\delta S_{A} has a physical meaning of the heat change of subsystem A during a process with fixed temperature TT. We can conclude from Eq. (17) that δQA(t)=H^A(t)\delta Q_{A}(t)=\langle\hat{H}_{A}(t)\rangle to the first order in F(t)F(t), this implies that for an open quantum system which is described by the canonical ensemble, if we know how its internal energy responds to a known perturbation, we then know how its EE responds to the same perturbation.

III.3 Exceptional states

Given an LRT of EE, a natural question arises: the sign of response is determined by the sign of driving, or the sign of α\alpha, consider a separable initial state, then SA(t0)=0S_{A}(t_{0})=0, we can always choose an appropriate sign of α\alpha such that the linear correction of EE is negative, that is δSA(t)<0\delta S_{A}(t)<0, then SA(t)=SA(t0)+δSA(t)<0S_{A}(t)=S_{A}(t_{0})+\delta S_{A}(t)<0, this is, of course, not true. We may have a similar question for maximally entangled states. Viewed from another perspective, the EE of product states and maximally entangled states must be extreme values, therefore δSA=0\delta S_{A}=0 for these states. To address these problems, we propose Theorem 1.

Theorem 1

For any product state or maximally entangled state of subsystems A and B which is an eigenstate of composite system A\cupB, if a time-dependent perturbation F(t)H^1F(t)\hat{H}_{1} (with F(t)1F(t)\ll 1) is applied to the composite system, then the linear response of EE of system A or B must be zero. Moreover, δSA/B(t)=𝒪(α2)\delta S_{A/B}(t)=\mathcal{O}(\alpha^{2}).

Proof. We here give the proof of the statement about separable states in Theorem 1, the proof about maximally states is similar and we show it in appendix C. The basic idea of this proof is to prove that lnρA(t0)\ln\rho_{A}(t_{0}) is diagonal while any diagonal element of δρA(t)\delta\rho_{A}(t) is zero, then trA[δρA(t)lnρA(t0)]=0\text{tr}_{A}[\delta\rho_{A}(t)\ln\rho_{A}(t_{0})]=0.

For a composite system consists of two subsystem A and B, if its eigenstates are separable or product states of A and B, then the Hamiltonian of this composite system must be an non-interacting Hamiltonian of A and B,

H^0=H^A+H^B\displaystyle\hat{H}_{0}=\hat{H}_{A}+\hat{H}_{B} (18)

before proceeding, we assume that H^0\hat{H}_{0} is non-degenerate, then both subsystem A and B are non-degenerate, their eigenstates and eigenvalues are given by

H^A|fiA\displaystyle\hat{H}_{A}|f_{i_{A}}\rangle =fiA|fiA\displaystyle=f_{i_{A}}|f_{i_{A}}\rangle (19)
H^B|gjB\displaystyle\hat{H}_{B}|g_{j_{B}}\rangle =gjB|gjB\displaystyle=g_{j_{B}}|g_{j_{B}}\rangle (20)

the eigenvectors of H^0\hat{H}_{0} are product states of the eigenvectors of A and B:

H^0(|fiA|gjB)=(fiA+gjB)|fiA|gjB\displaystyle\hat{H}_{0}\left(|f_{i_{A}}\rangle\otimes|g_{j_{B}}\rangle\right)=(f_{i_{A}}+g_{j_{B}})|f_{i_{A}}\rangle\otimes|g_{j_{B}}\rangle (21)

denoting |SiA,jB=|fiA|gjB|S_{i_{A},j_{B}}\rangle=|f_{i_{A}}\rangle\otimes|g_{j_{B}}\rangle, it is readily to see that if the initial state is diagonal with respect to basis {|SiA,jB}\{|S_{i_{A},j_{B}}\rangle\}, that is ρ0=iA,jBPiA,jB|SiA,jBSiA,jB|\rho_{0}=\sum_{i_{A},j_{B}}P_{i_{A},j_{B}}|S_{i_{A},j_{B}}\rangle\langle S_{i_{A},j_{B}}|, then the reduced density matrix of subsystem A is given by

ρA(t0)=trB[ρ0]=iA(jBPiA,jB)|fiAfiA|\displaystyle\rho_{A}(t_{0})=\text{tr}_{B}[\rho_{0}]=\sum_{i_{A}}\Big{(}\sum_{j_{B}}P_{i_{A},j_{B}}\Big{)}|f_{i_{A}}\rangle\langle f_{i_{A}}| (22)

which is diagonal with respect to eigenbasis of H^A\hat{H}_{A}, so is lnρA(t0)\ln\rho_{A}(t_{0}).

Given the intial state ρ0=iA,jBPiA,jB|SiA,jBSiA,jB|\rho_{0}=\sum_{i_{A},j_{B}}P_{i_{A},j_{B}}|S_{i_{A},j_{B}}\rangle\langle S_{i_{A},j_{B}}| which is diagonal, if [H^0,ρ0]=0[\hat{H}_{0},\rho_{0}]=0 then an arbitrary entry of the commutator in Eq.(11) is given by SiA,jB|[H^1,I(tτ),ρ0]|SkA,lB=(piA,jBpkA,lB)SiA,jB|H^1,I(tτ)|SkA,lB\langle S_{i_{A},j_{B}}|[\hat{H}_{1,I}(t-\tau),\rho_{0}]|S_{k_{A},l_{B}}\rangle=(p_{i_{A},j_{B}}-p_{k_{A},l_{B}})\langle S_{i_{A},j_{B}}|\hat{H}_{1,I}(t-\tau)|S_{k_{A},l_{B}}\rangle, and

δρ(t)m,m\displaystyle\langle\delta\rho(t)\rangle_{m,m} =it0t𝑑τSm|[H^1,I(tτ),ρ0]|SmF(τ)\displaystyle=-i\int_{t_{0}}^{t}d\tau\langle S_{m}|[\hat{H}_{1,I}(t-\tau),\rho_{0}]|S_{m}\rangle F(\tau)
+𝒪(F2)\displaystyle\ \ \ \ +\mathcal{O}(F^{2})
=𝒪(F2)\displaystyle=\mathcal{O}(F^{2}) (23)

where δρ(t)m,mρ(t)ρ0m,m\langle\delta\rho(t)\rangle_{m,m}\equiv\langle\rho(t)-\rho_{0}\rangle_{m,m}. The above equation implies that after applying perturbation, any diagonal element of linear correction to initial state is zero, does this means that any diagonal element of δρA(t)\delta\rho_{A}(t) is zero? Consider an arbitary off-diagonal element |SiA,jBSkA,lB||S_{i_{A},j_{B}}\rangle\langle S_{k_{A},l_{B}}| of the initial state,

trB(|SiA,jBSkA,lB|)=δjB,lB|fiAfkA|\displaystyle\text{tr}_{B}(|S_{i_{A},j_{B}}\rangle\langle S_{k_{A},l_{B}}|)=\delta_{j_{B},l_{B}}|f_{i_{A}}\rangle\langle f_{k_{A}}| (24)

which does not contribute to the diagonal element of subsystem A with respect to basis |fiA{|f_{i_{A}}}\rangle. As any diagonal element of ρ0\rho_{0} is zero, we may conclude that the δρA(t)\delta\rho_{A}(t) is hollow, that is the diagonal entries are all zero.

In summary, if the initial eigenstate is a separable state of A and B, moreover, if H^A+H^B\hat{H}_{A}+\hat{H}_{B} is non-degenerate, then lnρA(t0)\ln\rho_{A}(t_{0}) is diagonal, while δρA(t)\delta\rho_{A}(t) is hollow. Thus the linear response of EE δSA(t)=𝒪(F2)\delta S_{A}(t)=\mathcal{O}(F^{2}).

Refer to caption
Figure 1: XX spin chain with external driving on the first spin.
Refer to caption
Figure 2: Spectrum of linear response (yellow) and numerical results(blue) for driving amplitude α=0.1\alpha=0.1, other parameters: L=100L=100, J=2.0J=2.0, T=0.5πT=0.5\pi.
Refer to caption
Figure 3: Spectrum of linear response (yellow) and numerical results(blue) for driving amplitude α=2.5\alpha=2.5. Other parameters are the same as Fig. 2.

IV Linear Response of EE in XX Spin Chain

To verify our analytical result, we numerically investigate the dynamics of a one-dimension spin chain, which is composed of LL spin-1/21/2 particles coupled via XXXX-type interaction. Hamiltonian of this model in the absence of driving reads (we assume =1\hbar=1 hereafter)

H^0=λ2j=1Lσ^jz+J2j=1L1(σ^jxσ^j+1x+σ^jyσ^j+1y)\displaystyle\hat{H}_{0}=\frac{\lambda}{2}\sum_{j=1}^{L}\hat{\sigma}_{j}^{z}+\frac{J}{2}\sum_{j=1}^{L-1}(\hat{\sigma}_{j}^{x}\hat{\sigma}_{j+1}^{x}+\hat{\sigma}_{j}^{y}\hat{\sigma}_{j+1}^{y}) (25)

where σ^jk\hat{\sigma}_{j}^{k} (k=x,y,zk=x,\ y,\ z) are the Pauli operators with jj labeling the spins in the chain; λ\lambda denotes the longitudinal magnetic field exerted homogeneously on all the spins; JJ is the coupling strengths between the nearest-neighbor spins of the chain. Set the spin at the first site as subsystem A, a time-dependent perturbation of the following form is applied on this subsystem,

F(t)H^1(t)=F(t)σ^1z2\displaystyle F(t)\hat{H}_{1}(t)=F(t)\frac{\hat{\sigma}_{1}^{z}}{2} (26)

then the total Hamiltonian is given by H^(t)=H^0+H^source(t)\hat{H}(t)=\hat{H}_{0}+\hat{H}_{\text{source}}(t). Diagonalizing H^0\hat{H}_{0} in the single-excitation subspace, its eigenstates read

|φk=j=1Lei(2π/L)kjLσ^j+|{j}\displaystyle|\varphi_{k}\rangle=\sum_{j=1}^{L}\frac{e^{i(2\pi/L)kj}}{\sqrt{L}}\hat{\sigma}_{j}^{+}|\{\downarrow_{j}\}\rangle (27)

which is a spin wave with wave vector kk, and its corresponding eigenenergies are Ek=λ+2JcoskE_{k}=\lambda+2J\cos k. Here |{j}}|\{\downarrow_{j}\}\} is the ferromagnetic state of the chain with all its spin pointing to the e^z-\hat{e}_{z} direction and σ^j+=(σ^jx+iσ^jy)/2\hat{\sigma}_{j}^{+}=(\hat{\sigma}_{j}^{x}+i\hat{\sigma}_{j}^{y})/2. Any state of the system can be expressed as the superposition of these eigenstates

|ψ(t)=k=1Lck(t)|φk\displaystyle|\psi(t)\rangle=\sum_{k=1}^{L}c_{k}(t)|\varphi_{k}\rangle (28)

where ck(t)=φk|ψ(t)c_{k}(t)=\langle\varphi_{k}|\psi(t)\rangle. In numerical calculation, we choose ck(t)=δ1,kc_{k}(t)=\delta_{1,k} i.e. |ψ(t0)=|φ1|\psi(t_{0})\rangle=|\varphi_{1}\rangle, and F(t)=αcos(2πt/T)et2/2F(t)=\alpha\cos(2\pi t/T)e^{-t^{2}/2} with α\alpha a small prefactor and JT=πJT=\pi.

Through numerically solving the dynamics of XX spin chain in single-excitation subspace, we obtain δSA(t)\delta S_{A}(t), its spectrum δSA(ω)\delta S_{A}(\omega) is given by the discrete Fourier transform of numerical data. In Fig.2 and Fig.3, we show δSA(ω)\delta S_{A}(\omega) of the numerical result (blue line) and analytical one which only consider linear correction of dynamics (yellow line). There are two characteristic peaks in analytical result, to explain where do these peaks come from, we show the spectrum of RE(t)R_{E}(t) and F(t)F(t) in Fig.4, it is readily to see that the one near position 2ω/J±1.272\omega/J\approx\pm 1.27 result from the spectrum of response function χ(ω)\chi(\omega); the other peaks near 2ω/J±0.662\omega/J\approx\pm 0.66 result from the spectrum of driving F(ω)F(\omega). Comparing results with different α\alpha, it is found that when the driving is weak, i.e. α\alpha is small enough, numerical results match the linear response quite well, while α\alpha became large enough, there is a fairly obvious deviation between numerical spectrum and analytical one, especially at a position near 2ω/J=02\omega/J=0, 2ω/J=22\omega/J=2 and position near 2ω/J=2.52\omega/J=2.5, these deviations result from the emergent of non-linear response.

Refer to caption
Figure 4: The spectrum χ(ω)\chi(\omega) and F(ω)F(\omega), which is given by the discrete Fourier transform of response function RE(t)R_{E}(t) and driving F(t)F(t).

V Conclusion

An LRT of von Neumann entropy is developed for an open quantum system A which is subject to weak perturbation, we assumed A is a subsystem of composite system ABA\cup B, and the composite system is initially at a state which density matrix is commuted with the unperturbed Hamiltonian. We found that the LRT of von Neumann entropy to external perturbation is given by the linear response of an observable, using this observable we define the Kubo formula of EE which has the same properties of its conventional counterpart; using the LRT of EE, we demonstrate that there is no linear response of von Neumann entropy for a non-degenerate system which is initially at separable states and maximally entangled states. To verify our analytical results, we numerically solve the dynamics of EE for a subsystem of a XXXX spin chain. Numerical results show that when the perturbation is weak enough, the dynamics of EE will approach the result given by LRT. The LRT of EE provides a useful tool in understanding and investigating the dynamics of entanglement.

VI acknowledgments

We acknowledge helpful discussions with Han-Tao Lu and Chong Chen. This work was supported by the Natural Science Foundation of Guangdong Province (Grant No.2017B030308003), the Key R&D Program of Guangdong province (Grant No. 2018B030326001), the Science, Technology and Innovation Commission of Shenzhen Municipality (Grant No.JCYJ20170412152620376 and No.JCYJ20170817105046702 and No.KYTDPT20181011104202253), the National Natural Science Foundation of China (Grant No.11875160, No.U1801661 and No.11905100), the Economy, Trade and Information Commission of Shenzhen Municipality (Grant No.201901161512), the Guangdong Provincial Key Laboratory(Grant No.2019B121203002).

Appendix A Derivation of Eq. (11)

In interaction picture, we formally integral the Liouville equation and obtain

ρI(t)=ρ0it0t𝑑τ[F(τ)H^1,I(τ),ρI(τ)]\displaystyle\rho_{I}(t)=\rho_{0}-i\int_{t_{0}}^{t}d\tau[F(\tau)\hat{H}_{1,I}(\tau),\rho_{I}(\tau)] (29)

where ρ0ρ(t0)\rho_{0}\equiv\rho(t_{0}), H^1,I(t)=U^0(t,t0)H^1U^0(t,t0)\hat{H}_{1,I}(t)=\hat{U}_{0}^{\dagger}(t,t_{0})\hat{H}_{1}\hat{U}_{0}(t,t_{0}) and ρI(t)=U^0(t,t0)ρ(t)U^0(t,t0)\rho_{I}(t)=\hat{U}_{0}^{\dagger}(t,t_{0})\rho(t)\hat{U}_{0}(t,t_{0}), this expression is accurate. Assuming the value of F(t)F(t) is very small, we can expand ρI(t)\rho_{I}(t) in powers of F(t)F(t): ρI(t)=ρI(0)(τ)+FρI(1)(τ)+𝒪(F2)\rho_{I}(t)=\rho_{I}^{(0)}(\tau)+F\rho_{I}^{(1)}(\tau)+\mathcal{O}(F^{2}), inserting the expansion into (29), we obtain

ρI(t)=ρ0it0t𝑑τ[F(τ)H^1,I(τ),ρ0]+𝒪(F2),\displaystyle\rho_{I}(t)=\rho_{0}-i\int_{t_{0}}^{t}d\tau[F(\tau)\hat{H}_{1,I}(\tau),\rho_{0}]+\mathcal{O}(F^{2}), (30)

in the derivation of the above equation, we have used ρI(0)(t)=ρ(t0)ρ0\rho_{I}^{(0)}(t)=\rho(t_{0})\equiv\rho_{0}. we further assume the initial state ρ0\rho_{0} commutes with H^0\hat{H}_{0}, that is [ρ0,H^0]=0[\rho_{0},\hat{H}_{0}]=0. After transforming the density matrix in the above equation into Schrödinger picture, we have

δρ(t)=it0t𝑑τ[H^1,I(τt),ρ0]F(τ)+𝒪(F2)\displaystyle\delta\rho(t)=-i\int_{t_{0}}^{t}d\tau[\hat{H}_{1,I}(\tau-t),\rho_{0}]F(\tau)+\mathcal{O}(F^{2})

where δρ(t)=ρ(t)ρ(t0)\delta\rho(t)=\rho(t)-\rho(t_{0}), we thus get Eq. (11).

Appendix B Derivation of Eq. (13)

Working in the interaction picture, we insert Eq. (11) into Eq. (12), to the first order in F(t)F(t), we obtain the following equation

δSA(t)\displaystyle\delta S_{A}(t)
\displaystyle\approx it0t𝑑τF(τ)trA{s^AtrB([H^1,I(τt),ρ0])}\displaystyle-i\int_{t_{0}}^{t}d\tau F(\tau)\text{tr}_{A}\left\{\hat{s}_{A}\text{tr}_{B}\left(\big{[}\hat{H}_{1,I}(\tau-t),\rho_{0}\big{]}\right)\right\}
\displaystyle\approx it0t𝑑τF(τ)tr{s^A[H^1,I(τt),ρ0]}\displaystyle-i\int_{t_{0}}^{t}d\tau F(\tau)\text{tr}\left\{\hat{s}_{A}\big{[}\hat{H}_{1,I}(\tau-t),\rho_{0}\big{]}\right\} (31)

where s^AlnρA(t0)\hat{s}_{A}\equiv-\ln\rho_{A}(t_{0}). Using the cyclic property of trace tr(XY)=tr(YX)\text{tr}(XY)=\text{tr}(YX), we rewrite the above equation as

δSA(t)=it0t[s^A,H^1,I(τt)]0F(τ)𝑑τ+𝒪(F2)\displaystyle\delta S_{A}(t)=-i\int_{t_{0}}^{t}\langle\big{[}\hat{s}_{A},\hat{H}_{1,I}(\tau-t)\big{]}\rangle_{0}F(\tau)d\tau+\mathcal{O}(F^{2}) (32)

denoting RE(t,τ)=iθ(tτ)[s^A,H^1,I(τt)]0R_{E}(t,\tau)=-i\theta(t-\tau)\langle[\hat{s}_{A},\hat{H}_{1,I}(\tau-t)]\rangle_{0}, the above equation can be simplified into

δSA(t)t0RE(t,τ)F(τ)𝑑τ+𝒪(F2)\displaystyle\delta S_{A}(t)\approx\int_{t_{0}}^{\infty}R_{E}(t,\tau)F(\tau)d\tau+\mathcal{O}(F^{2}) (33)

we thus get Eq. (13).

It is worth noting that the Kubo formula RE(t,τ)R_{E}(t,\tau) is invariant under time translations, that is, RE(t,τ)=RE(tτ)R_{E}(t,\tau)=R_{E}(t-\tau).

Appendix C No linear response for maximally entangled states

Starting from Eq. (11), and assume [ρ0,H^0]=0[\rho_{0},\hat{H}_{0}]=0, we may obtain the change in the time evolution state of subsystem A:

δρA(t)\displaystyle\delta\rho_{A}(t) =trB[δρ(t)]\displaystyle=\text{tr}_{B}[\delta\rho(t)]
=it0t𝑑τtrB([H^1,I(τt),ρ0])F(τ)+𝒪(F2)\displaystyle=-i\int_{t_{0}}^{t}d\tau\text{tr}_{B}\left([\hat{H}_{1,I}(\tau-t),\rho_{0}]\right)F(\tau)+\mathcal{O}(F^{2}) (34)

inserting the above equation into δSA(t)=trA[δρA(t)lnρA(t)]\delta S_{A}(t)=-\text{tr}_{A}[\delta\rho_{A}(t)\ln\rho_{A}(t)], we have

δSA(t)\displaystyle\delta S_{A}(t)
=it0t𝑑τtrA{trB([H^1,I(τt),ρ0])s^A}+𝒪(F2).\displaystyle=-i\int_{t_{0}}^{t}d\tau\text{tr}_{A}\left\{\text{tr}_{B}\left([\hat{H}_{1,I}(\tau-t),\rho_{0}]\right)\hat{s}_{A}\right\}+\mathcal{O}(F^{2}). (35)

Consider a quantum system which consists of two subsystems A and B, the degree of freedom of A and B are both dd, assuming initially the total system is at maximally entangled state: ρ0=1di,j=1d|iAiBjAjB|\rho_{0}=\frac{1}{d}\sum_{i,j=1}^{d}|i_{A}i_{B}\rangle\langle j_{A}j_{B}|, where {|iA}\{|i_{A}\rangle\} and {|jB}\{|j_{B}\rangle\} are complete sets of orthonormal basis of subsystem A and B respectively, we then have

ρA(t0)=trB(ρ0)=1di|iAiA|=𝕀Ad\rho_{A}(t_{0})=\text{tr}_{B}\left(\rho_{0}\right)=\frac{1}{d}\sum_{i}|i_{A}\rangle\langle i_{A}|=\frac{\mathbb{I}_{A}}{d}

then lnρA(t0)=𝕀Alnd\ln\rho_{A}(t_{0})=-\mathbb{I}_{A}\ln d, which is diagonal. On the other hand,

trB([H^1,I(τt),ρ0])\displaystyle\text{tr}_{B}\left([\hat{H}_{1,I}(\tau-t),\rho_{0}]\right)
=\displaystyle= i,j=1d[H^1,I(jBiB)(τt)|iAjA|H^1,I(iBjB)(τt)|jAiA|]\displaystyle\sum_{i,j=1}^{d}\big{[}\hat{H}_{1,I}^{(j_{B}i_{B})}(\tau-t)|i_{A}\rangle\langle j_{A}|-\hat{H}_{1,I}^{(i_{B}j_{B})}(\tau-t)|j_{A}\rangle\langle i_{A}|\big{]} (36)

where H^1,I(jBiB)(τt)jB|H^1,I(τt)|iB\hat{H}_{1,I}^{(j_{B}i_{B})}(\tau-t)\equiv\langle j_{B}|\hat{H}_{1,I}(\tau-t)|i_{B}\rangle. It is not hard to see that the above operator is hollow with respect to basis {|iA}\{|i_{A}\rangle\}, therefore according to Eq.(36), we conclude that the response of EE to first order in α\alpha is zero: δSA(t)0\delta S_{A}(t)\approx 0.

References