Exact dynamics and thermalization of open quantum systems coupled to reservoir through particle exchanges
Abstract
In this paper, we study the exact dynamics of general open systems interacting with its environment through particle exchanges. The paper includes two main results. First, by taking advantage of the propagating function in the coherent state representation, we solve the exact master equation, whose solution is expressed in terms of the Keldysh nonequilibrium Green functions. Second, in the dynamical perspective, we provide a rigorous thermalization process of open quantum systems.
I Introduction
In the realistic world, physical systems are inevitably coupled to environments, which makes the theory of open quantum systems vastly used in many fields of physics, chemistry, biology, and engineering. The systematic study of open quantum systems has aroused the researchers’ interest since the 1960s Feynman and Vernon (1963); Schwinger (1961); Zwanzig (1960); Nakajima (1958); Breuer et al. (2002); Weiss (2012); Gardiner and Zoller (2004), and becomes more and more important for the prosperously developing field of quantum information processing Gardiner and Zoller (2004); Wiseman and Milburn (2009); Nielsen and Chuang (2002), quantum transport theory Haug and Jauho (2008); Yang and Zhang (2016), and rapidly improving time-resolved measurement technologies Wickenhauser et al. (2005); Kaldun et al. (2016). One of the most crucial problems in dealing with open quantum systems is how to determine explicitly the evolution of open quantum system states, through which all the information about the system dynamics can be obtained. However, due to the contamination of the huge environment, the system dynamics is non-unitary and always involves complicated fluctuation and dissipation. As a consequence, up to date, most open quantum systems can only be dealt with perturbative methods, such as Born-Markov approximation or cutting-off in Nakajima-Zwanzig operator projective technique Nakajima (1958); Zwanzig (1960); Breuer et al. (2002). Only a few classes of open quantum systems, e.g., the harmonic oscillator in quantum Brownian motion Caldeira and Leggett (1983); Hu et al. (1992) and the open systems interacting with particle-exchange coupling Zhang et al. (2012), can an exact master equation be obtained, let alone the exact state evolution.
In a series of papers Tu and Zhang (2008); Jin et al. (2010); Lei and Zhang (2012); Zhang et al. (2012), we have obtained the general exact master equation for the open quantum systems whose interaction only involves particle exchanges. This large class of open systems is an extension of the famous Fano-Anderson model Anderson (1961); Fano (1961) and characterizes many physical phenomena in different systems, such as Fano resonance in atomic and condensed matter physics Miroshnichenko et al. (2010), Anderson localization in many-body systems Anderson (1958), photon-atom bound states in photonic crystal Yablonovitch (1987); John (1987); Kofman et al. (1994), quantum transport in quantum dot junctions Haug and Jauho (2008), impurity defects in solids Mahan (2013), etc. Although it has been studied for decades, the exact general master equation has not been derived until very recently Tu and Zhang (2008); Jin et al. (2010); Lei and Zhang (2012); Zhang et al. (2012). With the progress, various perspectives of the open systems can be studied, e.g., memory effects Yang et al. (2013), entanglement dynamics Tan et al. (2011); Lin et al. (2016); An and Zhang (2007), decoherence Tu and Zhang (2008); Liu et al. (2016); Yang and Wu (2014), the fluctuation-dissipation theorem Zhang et al. (2012), exact transient quantum transport Yang and Zhang (2018); Lo et al. (2015); Yang and Zhang (2016), etc.
In this paper, one main result we obtained is the solution of the exact master equation, i.e., the exact form of the reduced density matrix at an arbitrary later time. Note that the time-dependence of the reduced density matrix includes complete information about the system dynamics, it offers more perspectives in studying the memory effects Li et al. (2018); Breuer et al. (2016); Rivas et al. (2014), entanglement dynamics and dynamical phase transition Heyl (2018) of such class of open systems. Using the result of the reduced density matrix, we also study the general thermalization in this paper, which is another important and hot topic for both experimentalists Trotzky et al. (2012); Gring et al. (2012) and theorists Deutsch (1991); Kosloff (2013); Ali et al. (2018); Xiong et al. (2015); Srednicki (1994); Linden et al. (2009); Short and Farrelly (2012); Reimann (2008); Rigol (2009); Polkovnikov et al. (2011); Cazalilla and Rigol (2010); Hsiang et al. (2018) in recent years. In equilibrium statistic mechanics, the thermal distribution is based on the assumption of equal probability of all permissible microstates Callen (1998). However, the foundation of thermalization has always been a tough problem and a long-term goal of physicists Polkovnikov et al. (2011). In the past decade, by taking advantage of the eigenstate thermalization hypothesis, researchers study the thermalization of closed quantum systems and make interesting contributions in understanding the physics beneath it Deutsch (1991); Kosloff (2013); Ali et al. (2018); Srednicki (1994); Linden et al. (2009); Short and Farrelly (2012); Reimann (2008); Rigol (2009); Polkovnikov et al. (2011); Cazalilla and Rigol (2010). In contrast, in this paper, we discuss the thermalization of open systems from the dynamical perspective. With the exact evolution of the open quantum systems, such an aim is achieved.
The rest of the paper is organized as follows. In Sec. II, we introduce the system we concerned about and briefly review the previous results related to our present work. In Sec. III, we shall derive the exact solution of the reduced density matrix. In Sec. IV, we explain the physical consequences of the solution and discuss the general thermalization of the open systems. A brief summary will be given in Sec. V.
II Overview of the exact master equation
The system we are interested are characterized by the Hamiltonian , and interacts with bosonic (fermionic) environment described by , via the interaction , where () is the creation (annihilation) operator of the th level in the system; () is the creation (annihilation) operator of the -mode in the environment; and characterize the energy levels of the system and the environment, and is the interaction strength between them. If the creation and annihilation operators satisfy the commutation (anti-commutation) relations, the corresponding systems is bosonic (fermionic).
To make the notations more compact, we denote that , , and the matrix constituted of as . The spectrum of the environment and the system-environment interaction can be specified with the spectral density function , whose elements are defined through . We assume that initially the system and the environment is decoupled Leggett et al. (1987), i.e., , where is an arbitrary physical state of the system and is the thermal state of the environment with inverse temperature and chemical potential . Here, denotes the grand partition function of the environment with the upper and lower sign corresponding to the bosonic and fermionic case respectively. Hereafter, we also use this convention.
The reduced density matrix at latter time is connected to the initial state through the propagating function in the coherent state representation, i.e.,
(1) |
where and , with the components being complex (Grassmannian) for bosons (fermions); () is the unnormalized coherent state with () standing for the vacuum state Zhang et al. (1990); is the integral measure, and for bosons and for fermions. With coherent state path-integral approach, the propagating function can be derived as Jin et al. (2010); Lei and Zhang (2012)
(2) |
where , , with the matrices and being respectively the abbreviations of the nonequilibrium Green functions and . More specifically, is the spectral Green function with elements defined as Kadanoff (2018), and is related to the system’s particle number originated from the environment, reading Zhang et al. (2012)
(3) |
Here, denotes a system-bath correction, with standing for the initial particle distribution of the environment.
Taking advantage of Eqs. (1)-(2), the master equation has been derived Tu and Zhang (2008); Jin et al. (2010); Lei and Zhang (2012); Zhang et al. (2012), reading
(4) |
In this master equation, is the renormalized system Hamiltonian; and characterize the dissipation and fluctuation rate induced by the system’s coupling with the environment, respectively. The coefficients , and are determined only by the non-equilibrium Green functions and Zhang et al. (2012).
Furthermore, the Green function satisfies the Dyson equation and the solution can be expressed as Zhang et al. (2012)
(5) |
The quantities , and are all connected with the Green function in the energy domain
(6) |
where is the self-energy correction. is the energy of the th localized mode and is determined by the singularities of , i.e., it satisfies the equation . is the Hermitian matrix characterizing the amplitude of the th localized mode, reading , where is the positively-oriented curve in the neighbourhood of . is the spectrum of the system broadened by the environment and reads 111The ’s and are the discrete and continuous components of the modified spectrum of the system , respectively. They are connected through the relation , where is defined through (See Ref. Zhang et al. (2012))..
III The exact system state evolution
With the initial system state , and the Green functions and , the reduced density matrix at arbitrary instant can be determined from Eq. (II) in principle. However, because the coefficients are all time-dependent, the problem is too complicated to solve directly. Luckily, one can solve it by taking advantage of the propagating function in Eqs. (1)-(2), as is shown in this section. First, in Sec. III.1, we shall introduce some conventions and derive ’s elements in the coherent state representation. In Sec. III.2, we obtain ’s elements at arbitrary later time. In Sec. III.3, the exact form of will be given.
III.1 Representation of the initial system state
To obtain the density matrix elements at arbitrary time, one needs to carry out the integrals of Eq. (1) with the explicit expression of . Define the Fock state
(7) |
where , , and denotes the particle number in the th level. For bosons, the takes the values , while for fermionic systems, is either or . Corresponding to the state , we define a class of sequences with the form
(8) |
Then the initial system density matrix can be generally expressed as
(9) |
where the summation is over all the possible physical pairs of and . That is, except for the constraints over and individually, for fermions and massive bosons, the sequences and together satisfy the constraint . With Eq. (9) and the definition of coherent states, it is easy to find
(10) |
where and .
III.2 Evolution of the density matrix elements in the coherent state representation
Following Eqs. (2) and (10), Eq. (1) can be re-expressed as
(11) |
In order to obtain the explicit form of , one needs to simplify the integral. The result is given in the following (See Appendix A for the details).
III.2.1 Bosonic case
In the bosonic case, the result is
(12) |
where we have used the following conventions:
-
(i)
is a subsequence of and is its complement, with for each . By subsequence, we mean deleting some elements in a sequence and keeping the remaining elements in the original order. The summation is over all the possible ways of choosing and in and , respectively, with each pair satisfying the constraint ;
- (ii)
-
(iii)
stands for the permanent of the matrix Tillmann et al. (2013);
-
(iv)
, .
III.2.2 Fermionic case
In the fermionic scenario, the integral can be simplified to
(13) |
Here, we have used the conventions:
-
(i)
, , , and the summation follow the same conventions as those in the bosonic case, except that all the elements in the sequences appear at most once;
-
(ii)
is the new sequence obtained by joining and end-to-end. when is an even/odd permutation of ;
-
(iii)
follows the same convention as in the bosonic case. However, as all the elements appear at most once in and , is in fact a submatrix of Tillmann et al. (2013);
-
(iv)
stands for the determinant of the matrix ;
-
(v)
and follow the same convention as in the bosonic case.
III.2.3 Common expression
III.3 Exact form of the density matrix
The reduced density matrix satisfying Eq. (17) can be explicitly written as
(18) |
where is a thermal-like state Xiong et al. (2015); and . Equation (18) is directly obtained from Eq. (17) with the identity
(19) |
which can be easily derived with the properties of coherent states that
(20a) | |||
(20b) |
and
(21) |
IV Physical interpretation of the solution and the thermalization
In this section, we shall discuss the physics beneath the solution. The physical interpretation of the solution will be given in Sec. IV.1, and the thermalization problem will be discussed in Sec. IV.2.
IV.1 Physical interpretation of the solution
To explain the physical consequence contained in Eq. (18), we consider two limiting cases first. One is that there is no particle in the environment initially, and the other is no particle in the system initially. Finally, we shall consider the joint effect and the general solution.
IV.1.1 No particle in the environment initially
In this case, the environment is initially in a vacuum state, so that and . Consequently, in Eq. (18), , , and reduces to , , and , respectively. As a result, is simply reduced to
(22) |
Compared with the form of the initial state of Eq. (9), one can see that the factor evolves to . Note that quantifies the particles maintained in the system and quantifies their loss into the environment. Therefore, the summation describes all the possibilities that part of the system particles maintains in the system while the others dissipate into the environment. That is, the solution in Eq. (22) precisely describes the pure dissipation process. The sign and the quantity are the manifestations of the particle exchange symmetry, which is quite similar to the cases encountered in the boson and fermion sampling Aaronson and Arkhipov (2011).
IV.1.2 No particle in the system initially
In this case, the system initial state reads , i.e., all the ’s are except the one that and are both empty sequence. Following Eq. (18), one can easily find that
(23) |
For understanding the physical consequence of Eq. (23), one needs to consider the physical meaning of . We introduce the spectral Green function of the total system, i.e., , where is the energy matrix of the total system Hamiltonian. Formally, can be written in matrix blocks, i.e.,
(26) |
where is the spectral Green function between X and Y, with S (E) being the abbreviation of the system (environment). Equivalent to Eq. (3), can be expressed as
(27) |
where is the particle distribution of the environment. From the equation, it is obvious that characterizes the average particle number transported from all the energy levels of the environment to the system. Therefore, is a thermal-like state completely contributed by the environmental particles propagating to the system, with the average particle number characterized by . In other words, the solution (23) comes from the thermal fluctuation process.
IV.1.3 Joint effect between fluctuation and dissipation
Now we consider the general result of Eq. (18). From Sec. IV.1.1, we can conclude that particles initially in the system contribute to the terms , , and . From Sec. IV.1.2, we know that originates from the particles initially in the environment. When initially particles coexist in both the system and environment, the dissipation property is modified due to the fluctuation. That is, and are modified as and , respectively. Due to the effect , the probability of the system particles dissipating into the environment becomes larger for bosons, while it becomes less for fermions. Correspondingly, the modification reveals that, for bosons, the amplitude of the system particles maintaining in the system becomes smaller; while for fermions, it becomes larger. These are due to the statistic properties of identical particles. For bosons, if an environmental level is occupied with some particles, the probability of the system particles hopping into that level becomes larger, as is described in Feynman lectures Feynman et al. (2011) and manifested in many phenomena such as superradience and Bose-Einstein condensation. While for fermions, due to the Pauli exclusion principle, if an environmental level is occupied, the transition onto it is forbidden.
IV.2 The thermalization
In this subsection, we shall study the asymptotic behavior of the state as the time approaches infinity. The absence or presence of localized modes determines whether the system can be finally thermalized, so we consider these two cases separately.
In the case that there are no localized modes (see Appendix B), and would finally evolve to
(28a) | |||
(28b) |
In this case, the final state of the system would be (See Appendix C for the details)
(29) |
Equation (29) implies that the final particle distribution in the system is completely characterized by the matrix Sharma and Rabani (2015), i.e.,
(30) |
With the expression of in Eq. (28b) and the properties of the spectral function (that it is positive-semidefinite and ), the final particle distribution can be seen as a weighted sum of the Bose/Fermi distribution. That is, without localized modes, the system would finally reach a thermal-like state, instead of the conventional thermal state.
When the coupling strength between the system and the environment is very weak, then the spectral density and the Lamb shift both tend to vanish, i.e.,
(31) |
Following
(32) |
and the careful analysis in Appendix D, one can find that under condition (31),
(33) |
That is, when the system-environment coupling becomes very weak, the broadening and Lamb shift of the system energy levels also vanishes, making the spectrum of the system converging to that of the isolated system. As a consequence of Eqs. (33) and (28b), approaches to the conventional Bose/Fermi distribution, i.e.,
(34) |
Thus, Eq. (29) converges to
(35) |
which is exactly the thermal state of the system in the grand canonical ensemble with inverse temperature and chemical potential of the environment. This provides a rigorous proof that in the weak-coupling limit, the exact evolution of an open quantum system would reproduce in the steady state limit the thermal state in conventional statistic mechanics.
On the other hand, if there are localized modes, their contribution to the oscillations in (See Eq. (5)) would survive as approaches infinity, i.e.,
(36) |
Following the expression of the reduced density matrix in Eq. (18), the final state must be expressed in terms of the coefficients , i.e., the system keeps the memory of its initial state. Therefore, the system cannot be thermalized.
V Summary
In this paper, we have investigated a general solution of open quantum systems interacting with the environment through particle exchanges. The exact evolution of the reduced density matrix is given in terms of the nonequilibrium Green functions. We explained the physical consequences of the solution. With the exact density matrix, we study the thermalization process. We obtain the result of equilibrium statistical mechanics from the dynamical perspective and go beyond it. That is, when there are no localized modes and the system-environment coupling is very weak, the final state would be as expected from the conventional statistical mechanics; for no localized modes but relatively strong coupling regime, the steady state would be thermal-like, which departures from the prediction of conventional statistical mechanics; when there are localized modes, the system keeps the memory of the initial state and can not be thermalized.
With the explicit expression of the reduced density matrix, one can obtain the complete information about the system dynamics, which is quite important for the rapidly developing quantum thermodynamics and quantum information, because their central physical quantity, entropy, is directly related to the state. It is also noteworthy that the model studied in our work involves non-Markovian nature. With the explicit form of the density matrix evolution, one can study the memory effects from more perspectives, e.g., quantum coherence, entanglement, and dynamical phase transition. Although we only consider the single-reservoir case, our result can be directly extended to the multi-reservoir case by just extending the corresponding expressions of the nonequilibrium Green functions and to multi-reservoirs (multi-leads in nano/quantum devices). Therefore, it is also easy to apply to quantum transport theory.
Acknowledgements.
We thank Yu-Wei Huang, Matisse Wei-Yuan Tu and Li Li for helpful discussions. This work is supported by the Ministry of Science and Technology of the Republic of China under the Contracts No. MOST 107-2811-M-006-534 and No. MOST 108-2811-M-006-518.Appendix A Simplification of the integral
A.1 Bosonic case
From Eq. (11), for bosons,
(48) | ||||
(60) | ||||
(61) |
where we have used the convention and the formula of Gaussian integral Kamenev and Levchenko (2009).
is only related to the coefficient of in the polynomial expansion of . Note
(62) |
The terms with factor can be obtained through that
-
(i)
contributes to ,
-
(ii)
contributes to ,
-
(iii)
contributes to ,
where satisfy the constraint
(63a) | |||
(63b) | |||
(63c) | |||
(63d) |
Note,
-
(i)
the coefficient of in is ;
-
(ii)
the coefficient of in is ;
-
(iii)
the coefficient of
in is .
Therefore, the coefficient of in is , and the total coefficient of in Eq. (62) is the summation of all these terms with ’s ’s satisfying Eq. (63), i.e.,
(64) |
Also note that
(65) |
therefore
(66) |
where stands for the binomial coefficient. The factor is the number of ways of obtaining from as well as obtaining from . So we can transform the summation over all the possible ’s and ’s to the summation of all the possible ways of obtaining subsequences ’s and ’s from and , respectively. Therefore, the factor is absorbed in the sum and the result can be reexpressed as
(67) |
A.2 Fermionic case
From Eq. (11), for fermions, we have
(79) | ||||
(91) | ||||
(92) |
where we have used the convention and the formula of Grassmannian Gaussian integral Kamenev and Levchenko (2009).
is only related to the coefficient of in the polynomial expansion of . Note
(93) |
the terms with factor can be obtained through that
-
(i)
contributes to ,
-
(ii)
contributes to ,
-
(iii)
contributes to ,
where satisfy the constraint
(94a) | |||
(94b) | |||
(94c) | |||
(94d) |
Note that
-
(i)
the coefficient of in is ;
-
(ii)
the coefficient of in is ;
-
(iii)
the coefficient of in is ;
and
(95) |
therefore, the coefficient of in is . The total coefficient of in Eq. (93) is the summation of all the possible terms with ’s ’s satisfying Eq. (94), which is denoted as
(96) |
Also note that for Grassmannian variables
(97) |
therefore
(98) |
We can express the result in a new way, i.e., as the summation over all the possible ways of obtaining ’s and ’s from and , respectively. Because the way obtaining and from and in the fermionic case is unique, can be directly replaced by . Therefore, the final result can be formulated as
(99) |
Appendix B Asymptotic form of as approaches infinity
Except for the formula in Ref. Zhang et al. (2012), the elements of can also be expressed in terms of the system-environment Green function, reading
(100) |
where and . In order to obtain in the long-time limit, we need to find the asymptotic behavior of .
Note that
(101) |
and
(102) |
can be simplified to
(103) |
When there are no localized modes in the total system, there is no singularity in and there is only a pole located at in the integrated function of Eq. (103). Using the contour integral, one obtains
(104) |
in the long-time limit therefore reads
(105) |
In terms of the matrix representation, it can be expressed as
(106) |
Following Eq. (114), it can be further simplified to
(107) |
This is the standard equilibrium fluctuation-dissipation theorem.
Appendix C Equilibrium state of the system
If the total system possesses no localized modes, the open system would finally reach an equilibrium state. Following the expression of the reduced density matrix in Eq. (18) and the asymptotic expression of and in Eq. (28), when approaches infinity, only the terms with would survive, i.e.,
(108) |
Because reduces to the identity matrix as vanishes, the matrix would approach to a block-diagonal matrix in the form
(109) |
where
(110) |
The permanent and determinant of then share a common expression, reading
(111) |
Following Eq. (108), can be simplified to
(112) |
With the normalization condition that , it can be finally written as
(113) |
Appendix D Asymptotic form of as the coupling strength vanishes
When and , the spectrum
(114) |
is vanishing for the values of that the matrices are invertible. Because and , the condition can be simplified as that is invertible. Therefore, is nonvanishing only for equaling an eigenvalue of . So in order to grasp the asymptotic behavior of , one only needs to analyze the behavior of around .
Consider the behavior of for . Denote the eigenspace of corresponding to the eigenvalue as , then the matrix can be written in blocks, reading
(115) |
where the subscripts and are corresponding to the space and its orthogonal complement, respectively. Because and , the inverse of is approximately
(116) |
where is the identity in . (In this equation, we have kept the term for could also be small.) Following Eqs. (114) and (116), and using the properties that and , the expression of around can be approximately written as
(117) |
When and approach to , the real part of is dominant by , so the above equation can be further simplified to
(118) |
After expressing the right hand side of the equation in the eigenbasis of , one can easily find that the diagonal elements all approach to as vanishes. Consequently, for near ,
(119) |
For every , such conclusion is always true. Therefore, for all ,
(120) |
References
- Feynman and Vernon (1963) R. Feynman and F. Vernon, Ann. Phys. 24, 118 (1963).
- Schwinger (1961) J. Schwinger, J. Math. Phys. 2, 407 (1961).
- Zwanzig (1960) R. Zwanzig, Int. J. Chem. Phys. 33, 1338 (1960).
- Nakajima (1958) S. Nakajima, Prog. Theor. Phys. 20, 948 (1958).
- Breuer et al. (2002) H.-P. Breuer, F. Petruccione, et al., The Theory of Open Quantum Systems (Oxford University Press on Demand, 2002).
- Weiss (2012) U. Weiss, Quantum Dissipative Systems, Vol. 13 (World scientific, 2012).
- Gardiner and Zoller (2004) C. Gardiner and P. Zoller, Quantum Noise: A Handbook of Markovian and Non-Markovian Quantum Stochastic Methods with Applications to Quantum Optics, Vol. 56 (Springer Science & Business Media, 2004).
- Wiseman and Milburn (2009) H. M. Wiseman and G. J. Milburn, Quantum Measurement and Control (Cambridge university press, 2009).
- Nielsen and Chuang (2002) M. A. Nielsen and I. Chuang, “Quantum computation and quantum information,” (2002).
- Haug and Jauho (2008) H. Haug and A.-P. Jauho, Quantum Kinetics in Transport and Optics of Semiconductors, Vol. 2 (Springer, 2008).
- Yang and Zhang (2016) P.-Y. Yang and W.-M. Zhang, Front. Phys. 12, 127204 (2016).
- Wickenhauser et al. (2005) M. Wickenhauser, J. Burgdörfer, F. Krausz, and M. Drescher, Phys. Rev. Lett. 94, 023002 (2005).
- Kaldun et al. (2016) A. Kaldun, A. Blättermann, V. Stooß, S. Donsa, H. Wei, R. Pazourek, S. Nagele, C. Ott, C. Lin, J. Burgdörfer, et al., Science 354, 738 (2016).
- Caldeira and Leggett (1983) A. Caldeira and A. Leggett, Physica A 121, 587 (1983).
- Hu et al. (1992) B. L. Hu, J. P. Paz, and Y. Zhang, Phys. Rev. D 45, 2843 (1992).
- Zhang et al. (2012) W.-M. Zhang, P.-Y. Lo, H.-N. Xiong, M. W.-Y. Tu, and F. Nori, Phys. Rev. Lett. 109, 170402 (2012).
- Tu and Zhang (2008) M. W. Y. Tu and W.-M. Zhang, Phys. Rev. B 78, 235311 (2008).
- Jin et al. (2010) J. Jin, M. W.-Y. Tu, W.-M. Zhang, and Y. Yan, New J. Phys. 12, 083013 (2010).
- Lei and Zhang (2012) C. U. Lei and W.-M. Zhang, Ann. Phys. 327, 1408 (2012).
- Anderson (1961) P. W. Anderson, Phys. Rev. 124, 41 (1961).
- Fano (1961) U. Fano, Phys. Rev. 124, 1866 (1961).
- Miroshnichenko et al. (2010) A. E. Miroshnichenko, S. Flach, and Y. S. Kivshar, Rev. Mod. Phys. 82, 2257 (2010).
- Anderson (1958) P. W. Anderson, Phys. Rev. 109, 1492 (1958).
- Yablonovitch (1987) E. Yablonovitch, Phys. Rev. Lett. 58, 2059 (1987).
- John (1987) S. John, Phys. Rev. Lett. 58, 2486 (1987).
- Kofman et al. (1994) A. Kofman, G. Kurizki, and B. Sherman, J. Mod. Opt. 41, 353 (1994).
- Mahan (2013) G. D. Mahan, Many-Particle Physics (Springer Science & Business Media, 2013).
- Yang et al. (2013) W. L. Yang, J.-H. An, C. Zhang, M. Feng, and C. H. Oh, Phys. Rev. A 87, 022312 (2013).
- Tan et al. (2011) H.-T. Tan, W.-M. Zhang, and G.-X. Li, Phys. Rev. A 83, 062310 (2011).
- Lin et al. (2016) Y.-C. Lin, P.-Y. Yang, and W.-M. Zhang, Sci. Rep. 6, 34804 (2016).
- An and Zhang (2007) J.-H. An and W.-M. Zhang, Phys. Rev. A 76, 042127 (2007).
- Liu et al. (2016) J.-H. Liu, M. W.-Y. Tu, and W.-M. Zhang, Phys. Rev. B 94, 045403 (2016).
- Yang and Wu (2014) M.-J. Yang and S.-T. Wu, Phys. Rev. A 89, 022301 (2014).
- Yang and Zhang (2018) P.-Y. Yang and W.-M. Zhang, Phys. Rev. B 97, 054301 (2018).
- Lo et al. (2015) P.-Y. Lo, H.-N. Xiong, and W.-M. Zhang, Sci. Rep. 5, 9423 (2015).
- Li et al. (2018) L. Li, M. J. Hall, and H. M. Wiseman, Phys. Rep. 759, 1 (2018).
- Breuer et al. (2016) H.-P. Breuer, E.-M. Laine, J. Piilo, and B. Vacchini, Rev. Mod. Phys. 88, 021002 (2016).
- Rivas et al. (2014) Á. Rivas, S. F. Huelga, and M. B. Plenio, Rep. Prog. Phys. 77, 094001 (2014).
- Heyl (2018) M. Heyl, Rep. Prog. Phys. 81, 054001 (2018).
- Trotzky et al. (2012) S. Trotzky, Y.-A. Chen, A. Flesch, I. P. McCulloch, U. Schollwöck, J. Eisert, and I. Bloch, Nat. Phys. 8, 325 (2012).
- Gring et al. (2012) M. Gring, M. Kuhnert, T. Langen, T. Kitagawa, B. Rauer, M. Schreitl, I. Mazets, D. A. Smith, E. Demler, and J. Schmiedmayer, Science 337, 1318 (2012).
- Deutsch (1991) J. M. Deutsch, Phys. Rev. A 43, 2046 (1991).
- Kosloff (2013) R. Kosloff, Entropy 15, 2100 (2013).
- Ali et al. (2018) M. Ali, W.-M. Huang, W.-M. Zhang, et al., arXiv preprint arXiv:1803.04658 (2018).
- Xiong et al. (2015) H.-N. Xiong, P.-Y. Lo, W.-M. Zhang, F. Nori, et al., Sci. Rep. 5, 13353 (2015).
- Srednicki (1994) M. Srednicki, Phys. Rev. E 50, 888 (1994).
- Linden et al. (2009) N. Linden, S. Popescu, A. J. Short, and A. Winter, Phys. Rev. E 79, 061103 (2009).
- Short and Farrelly (2012) A. J. Short and T. C. Farrelly, New J. Phys. 14, 013063 (2012).
- Reimann (2008) P. Reimann, Phys. Rev. Lett. 101, 190403 (2008).
- Rigol (2009) M. Rigol, Phys. Rev. Lett. 103, 100403 (2009).
- Polkovnikov et al. (2011) A. Polkovnikov, K. Sengupta, A. Silva, and M. Vengalattore, Rev. Mod. Phys. 83, 863 (2011).
- Cazalilla and Rigol (2010) M. Cazalilla and M. Rigol, New J. Phys. 12, 055006 (2010).
- Hsiang et al. (2018) J.-T. Hsiang, C. H. Chou, Y. Subaşı, and B. L. Hu, Phys. Rev. E 97, 012135 (2018).
- Callen (1998) H. B. Callen, “Thermodynamics and an introduction to thermostatistics,” (1998).
- Leggett et al. (1987) A. J. Leggett, S. Chakravarty, A. T. Dorsey, M. P. A. Fisher, A. Garg, and W. Zwerger, Rev. Mod. Phys. 59, 1 (1987).
- Zhang et al. (1990) W.-M. Zhang, D. H. Feng, and R. Gilmore, Rev. Mod. Phys. 62, 867 (1990).
- Kadanoff (2018) L. P. Kadanoff, Quantum Statistical Mechanics (CRC Press, 2018).
- Tillmann et al. (2013) M. Tillmann, B. Dakić, R. Heilmann, S. Nolte, A. Szameit, and P. Walther, Nat. Photon. 7, 540 (2013).
- Aaronson and Arkhipov (2011) S. Aaronson and A. Arkhipov, in Proceedings of the Forty-Third Annual ACM Symposium on Theory of Computing (ACM, 2011) pp. 333–342.
- Feynman et al. (2011) R. P. Feynman, R. B. Leighton, and M. Sands, The Feynman Lectures on Physics, Vol. 3 (Basic books, 2011).
- Sharma and Rabani (2015) A. Sharma and E. Rabani, Phys. Rev. B 91, 085121 (2015).
- Kamenev and Levchenko (2009) A. Kamenev and A. Levchenko, Adv. Phys. (2009).