Quantum thermodynamics

Last updated

Quantum thermodynamics [1] [2] is the study of the relations between two independent physical theories: thermodynamics and quantum mechanics. The two independent theories address the physical phenomena of light and matter. In 1905, Albert Einstein argued that the requirement of consistency between thermodynamics and electromagnetism [3] leads to the conclusion that light is quantized, obtaining the relation . This paper is the dawn of quantum theory. In a few decades quantum theory became established with an independent set of rules. [4] Currently quantum thermodynamics addresses the emergence of thermodynamic laws from quantum mechanics. It differs from quantum statistical mechanics in the emphasis on dynamical processes out of equilibrium. In addition, there is a quest for the theory to be relevant for a single individual quantum system.

Contents

Dynamical view

There is an intimate connection of quantum thermodynamics with the theory of open quantum systems. [5] Quantum mechanics inserts dynamics into thermodynamics, giving a sound foundation to finite-time-thermodynamics. The main assumption is that the entire world is a large closed system, and therefore, time evolution is governed by a unitary transformation generated by a global Hamiltonian. For the combined system bath scenario, the global Hamiltonian can be decomposed into:

where is the system Hamiltonian, is the bath Hamiltonian and is the system-bath interaction. The state of the system is obtained from a partial trace over the combined system and bath: . Reduced dynamics is an equivalent description of the system dynamics utilizing only system operators. Assuming Markov property for the dynamics the basic equation of motion for an open quantum system is the Lindblad equation (GKLS): [6] [7]

is a (Hermitian) Hamiltonian part and :

is the dissipative part describing implicitly through system operators the influence of the bath on the system. The Markov property imposes that the system and bath are uncorrelated at all times . The L-GKS equation is unidirectional and leads any initial state to a steady state solution which is an invariant of the equation of motion . [5]

The Heisenberg picture supplies a direct link to quantum thermodynamic observables. The dynamics of a system observable represented by the operator, , has the form:

where the possibility that the operator, is explicitly time-dependent, is included.

Emergence of time derivative of first law of thermodynamics

When the first law of thermodynamics emerges:

where power is interpreted as and the heat current . [8] [9] [10]

Additional conditions have to be imposed on the dissipator to be consistent with thermodynamics. First the invariant should become an equilibrium Gibbs state. This implies that the dissipator should commute with the unitary part generated by . [5] In addition an equilibrium state is stationary and stable. This assumption is used to derive the Kubo-Martin-Schwinger stability criterion for thermal equilibrium i.e. KMS state.

A unique and consistent approach is obtained by deriving the generator, , in the weak system bath coupling limit. [11] In this limit, the interaction energy can be neglected. This approach represents a thermodynamic idealization: it allows energy transfer, while keeping a tensor product separation between the system and bath, i.e., a quantum version of an isothermal partition.

Markovian behavior involves a rather complicated cooperation between system and bath dynamics. This means that in phenomenological treatments, one cannot combine arbitrary system Hamiltonians, , with a given L-GKS generator. This observation is particularly important in the context of quantum thermodynamics, where it is tempting to study Markovian dynamics with an arbitrary control Hamiltonian. Erroneous derivations of the quantum master equation can easily lead to a violation of the laws of thermodynamics.

An external perturbation modifying the Hamiltonian of the system will also modify the heat flow. As a result, the L-GKS generator has to be renormalized. For a slow change, one can adopt the adiabatic approach and use the instantaneous system’s Hamiltonian to derive . An important class of problems in quantum thermodynamics is periodically driven systems. Periodic quantum heat engines and power-driven refrigerators fall into this class.

A reexamination of the time-dependent heat current expression using quantum transport techniques has been proposed. [12]

A derivation of consistent dynamics beyond the weak coupling limit has been suggested. [13]

Phenomenological formulations of irreversible quantum dynamics consistent with the second law and implementing the geometric idea of "steepest entropy ascent" or "gradient flow" have been suggested to model relaxation and strong coupling. [14] [15]

Emergence of the second law

The second law of thermodynamics is a statement on the irreversibility of dynamics or, the breakup of time reversal symmetry (T-symmetry). This should be consistent with the empirical direct definition: heat will flow spontaneously from a hot source to a cold sink.

From a static viewpoint, for a closed quantum system, the 2nd law of thermodynamics is a consequence of the unitary evolution. [16] In this approach, one accounts for the entropy change before and after a change in the entire system. A dynamical viewpoint is based on local accounting for the entropy changes in the subsystems and the entropy generated in the baths.

Entropy

In thermodynamics, entropy is related to the amount of energy of a system that can be converted into mechanical work in a concrete process. [17] In quantum mechanics, this translates to the ability to measure and manipulate the system based on the information gathered by measurement. An example is the case of Maxwell’s demon, which has been resolved by Leó Szilárd. [18] [19] [20]

The entropy of an observable is associated with the complete projective measurement of an observable, where the operator has a spectral decomposition:

where are the projection operators of the eigenvalue

The probability of outcome is The entropy associated with the observable is the Shannon entropy with respect to the possible outcomes:

The most significant observable in thermodynamics is the energy represented by the Hamiltonian operator and its associated energy entropy, [21]

John von Neumann suggested to single out the most informative observable to characterize the entropy of the system. This invariant is obtained by minimizing the entropy with respect to all possible observables. The most informative observable operator commutes with the state of the system. The entropy of this observable is termed the Von Neumann entropy and is equal to

As a consequence, for all observables. At thermal equilibrium the energy entropy is equal to the von Neumann entropy:

is invariant to a unitary transformation changing the state. The Von Neumann entropy is additive only for a system state that is composed of a tensor product of its subsystems:

Clausius version of the II-law

No process is possible whose sole result is the transfer of heat from a body of lower temperature to a body of higher temperature.

This statement for N-coupled heat baths in steady state becomes

A dynamical version of the II-law can be proven, based on Spohn's inequality: [22]

which is valid for any L-GKS generator, with a stationary state, . [5]

Consistency with thermodynamics can be employed to verify quantum dynamical models of transport. For example, local models for networks where local L-GKS equations are connected through weak links have been thought to violate the second law of thermodynamics. [23] In 2018 has been shown that, by correctly taking into account all work and energy contributions in the full system, local master equations are fully coherent with the second law of thermodynamics [24]

Quantum and thermodynamic adiabatic conditions and quantum friction

Thermodynamic adiabatic processes have no entropy change. Typically, an external control modifies the state. A quantum version of an adiabatic process can be modeled by an externally controlled time dependent Hamiltonian . If the system is isolated, the dynamics are unitary, and therefore, is a constant. A quantum adiabatic process is defined by the energy entropy being constant. The quantum adiabatic condition is therefore equivalent to no net change in the population of the instantaneous energy levels. This implies that the Hamiltonian should commute with itself at different times: .

When the adiabatic conditions are not fulfilled, additional work is required to reach the final control value. For an isolated system, this work is recoverable, since the dynamics is unitary and can be reversed. In this case, quantum friction can be suppressed using shortcuts to adiabaticity as demonstrated in the laboratory using a unitary Fermi gas in a time-dependent trap. [25] The coherence stored in the off-diagonal elements of the density operator carry the required information to recover the extra energy cost and reverse the dynamics. Typically, this energy is not recoverable, due to interaction with a bath that causes energy dephasing. The bath, in this case, acts like a measuring apparatus of energy. This lost energy is the quantum version of friction. [26] [27]

Emergence of the dynamical version of the third law of thermodynamics

There are seemingly two independent formulations of the third law of thermodynamics both originally were stated by Walther Nernst. The first formulation is known as the Nernst heat theorem, and can be phrased as:

The second formulation is dynamical, known as the unattainability principle [28]

At steady state the second law of thermodynamics implies that the total entropy production is non-negative. When the cold bath approaches the absolute zero temperature, it is necessary to eliminate the entropy production divergence at the cold side when , therefore

For the fulfillment of the second law depends on the entropy production of the other baths, which should compensate for the negative entropy production of the cold bath. The first formulation of the third law modifies this restriction. Instead of the third law imposes , guaranteeing that at absolute zero the entropy production at the cold bath is zero: . This requirement leads to the scaling condition of the heat current .

The second formulation, known as the unattainability principle can be rephrased as; [29]

The dynamics of the cooling process is governed by the equation:

where is the heat capacity of the bath. Taking and with , we can quantify this formulation by evaluating the characteristic exponent of the cooling process,

This equation introduces the relation between the characteristic exponents and . When then the bath is cooled to zero temperature in a finite time, which implies a violation of the third law. It is apparent from the last equation, that the unattainability principle is more restrictive than the Nernst heat theorem.

Typicality as a source of emergence of thermodynamic phenomena

The basic idea of quantum typicality is that the vast majority of all pure states featuring a common expectation value of some generic observable at a given time will yield very similar expectation values of the same observable at any later time. This is meant to apply to Schrödinger type dynamics in high dimensional Hilbert spaces. As a consequence individual dynamics of expectation values are then typically well described by the ensemble average. [30]

Quantum ergodic theorem originated by John von Neumann is a strong result arising from the mere mathematical structure of quantum mechanics. The QET is a precise formulation of termed normal typicality, i.e. the statement that, for typical large systems, every initial wave function from an energy shell is ‘normal’: it evolves in such a way that for most t, is macroscopically equivalent to the micro-canonical density matrix. [31]

Resource theory

The second law of thermodynamics can be interpreted as quantifying state transformations which are statistically unlikely so that they become effectively forbidden. The second law typically applies to systems composed of many particles interacting; Quantum thermodynamics resource theory is a formulation of thermodynamics in the regime where it can be applied to a small number of particles interacting with a heat bath. For processes which are cyclic or very close to cyclic, the second law for microscopic systems takes on a very different form than it does at the macroscopic scale, imposing not just one constraint on what state transformations are possible, but an entire family of constraints. These second laws are not only relevant for small systems, but also apply to individual macroscopic systems interacting via long-range interactions, which only satisfy the ordinary second law on average. By making precise the definition of thermal operations, the laws of thermodynamics take on a form with the first law defining the class of thermal operations, the zeroth law emerging as a unique condition ensuring the theory is nontrivial, and the remaining laws being a monotonicity property of generalised free energies. [32] [33]

Engineered reservoirs

Nanoscale allows for the preparation of quantum systems in physical states without classical analogs. There, complex out-of-equilibrium scenarios may be produced by the initial preparation of either the working substance or the reservoirs of quantum particles, the latter dubbed as "engineered reservoirs". There are different forms of engineered reservoirs. Some of them involve subtle quantum coherence or correlation effects, [34] [35] [36] while others rely solely on nonthermal classical probability distribution functions. [37] [38] [39] [40] Interesting phenomena may emerge from the use of engineered reservoirs such as efficiencies greater than the Otto limit, [36] violations of Clausius inequalities, [41] or simultaneous extraction of heat and work from the reservoirs. [35]

See also

Related Research Articles

<span class="mw-page-title-main">Quantum entanglement</span> Correlation between quantum systems

Quantum entanglement is the phenomenon that occurs when a duet of particles are generated, interact, or share spatial proximity in such a way that the quantum state of each particle of the group cannot be described independently of the state of the others, including when the particles are separated by a large distance. The topic of quantum entanglement is at the heart of the disparity between classical and quantum physics: entanglement is a primary feature of quantum mechanics not present in classical mechanics.

In quantum mechanics, a density matrix is a matrix that describes the quantum state of a physical system. It allows for the calculation of the probabilities of the outcomes of any measurement performed upon this system, using the Born rule. It is a generalization of the more usual state vectors or wavefunctions: while those can only represent pure states, density matrices can also represent mixed states. Mixed states arise in quantum mechanics in two different situations:

  1. when the preparation of the system is not fully known, and thus one must deal with a statistical ensemble of possible preparations, and
  2. when one wants to describe a physical system that is entangled with another, without describing their combined state; this case is typical for a system interacting with some environment.
<span class="mw-page-title-main">Quantum decoherence</span> Loss of quantum coherence

Quantum decoherence is the loss of quantum coherence, the process in which a system's behaviour changes from that which can be explained by quantum mechanics to that which can be explained by classical mechanics. In quantum mechanics, particles such as electrons are described by a wave function, a mathematical representation of the quantum state of a system; a probabilistic interpretation of the wave function is used to explain various quantum effects. As long as there exists a definite phase relation between different states, the system is said to be coherent. A definite phase relationship is necessary to perform quantum computing on quantum information encoded in quantum states. Coherence is preserved under the laws of quantum physics.

In physics, a partition function describes the statistical properties of a system in thermodynamic equilibrium. Partition functions are functions of the thermodynamic state variables, such as the temperature and volume. Most of the aggregate thermodynamic variables of the system, such as the total energy, free energy, entropy, and pressure, can be expressed in terms of the partition function or its derivatives. The partition function is dimensionless.

<span class="mw-page-title-main">Black hole thermodynamics</span> Area of study

In physics, black hole thermodynamics is the area of study that seeks to reconcile the laws of thermodynamics with the existence of black hole event horizons. As the study of the statistical mechanics of black-body radiation led to the development of the theory of quantum mechanics, the effort to understand the statistical mechanics of black holes has had a deep impact upon the understanding of quantum gravity, leading to the formulation of the holographic principle.

In quantum mechanics, the Gorini–Kossakowski–Sudarshan–Lindblad equation, master equation in Lindblad form, quantum Liouvillian, or Lindbladian is one of the general forms of Markovian master equations describing open quantum systems. It generalizes the Schrödinger equation to open quantum systems; that is, systems in contacts with their surroundings. The resulting dynamics is no longer unitary, but still satisfies the property of being trace-preserving and completely positive for any initial condition.

In quantum physics, a measurement is the testing or manipulation of a physical system to yield a numerical result. A fundamental feature of quantum theory is that the predictions it makes are probabilistic. The procedure for finding a probability involves combining a quantum state, which mathematically describes a quantum system, with a mathematical representation of the measurement to be performed on that system. The formula for this calculation is known as the Born rule. For example, a quantum particle like an electron can be described by a quantum state that associates to each point in space a complex number called a probability amplitude. Applying the Born rule to these amplitudes gives the probabilities that the electron will be found in one region or another when an experiment is performed to locate it. This is the best the theory can do; it cannot say for certain where the electron will be found. The same quantum state can also be used to make a prediction of how the electron will be moving, if an experiment is performed to measure its momentum instead of its position. The uncertainty principle implies that, whatever the quantum state, the range of predictions for the electron's position and the range of predictions for its momentum cannot both be narrow. Some quantum states imply a near-certain prediction of the result of a position measurement, but the result of a momentum measurement will be highly unpredictable, and vice versa. Furthermore, the fact that nature violates the statistical conditions known as Bell inequalities indicates that the unpredictability of quantum measurement results cannot be explained away as due to ignorance about "local hidden variables" within quantum systems.

The Unruh effect is a theoretical prediction in quantum field theory that states that an observer who is uniformly accelerating through empty space will perceive a thermal bath. This means that even in the absence of any external heat sources, an accelerating observer will detect particles and experience a temperature. In contrast, an inertial observer in the same region of spacetime would observe no temperature.

In physics, an open quantum system is a quantum-mechanical system that interacts with an external quantum system, which is known as the environment or a bath. In general, these interactions significantly change the dynamics of the system and result in quantum dissipation, such that the information contained in the system is lost to its environment. Because no quantum system is completely isolated from its surroundings, it is important to develop a theoretical framework for treating these interactions in order to obtain an accurate understanding of quantum systems.

In quantum mechanics, einselections, short for "environment-induced superselection", is a name coined by Wojciech H. Zurek for a process which is claimed to explain the appearance of wavefunction collapse and the emergence of classical descriptions of reality from quantum descriptions. In this approach, classicality is described as an emergent property induced in open quantum systems by their environments. Due to the interaction with the environment, the vast majority of states in the Hilbert space of a quantum open system become highly unstable due to entangling interaction with the environment, which in effect monitors selected observables of the system. After a decoherence time, which for macroscopic objects is typically many orders of magnitude shorter than any other dynamical timescale, a generic quantum state decays into an uncertain state which can be expressed as a mixture of simple pointer states. In this way the environment induces effective superselection rules. Thus, einselection precludes stable existence of pure superpositions of pointer states. These 'pointer states' are stable despite environmental interaction. The einselected states lack coherence, and therefore do not exhibit the quantum behaviours of entanglement and superposition.

In statistical mechanics, the microcanonical ensemble is a statistical ensemble that represents the possible states of a mechanical system whose total energy is exactly specified. The system is assumed to be isolated in the sense that it cannot exchange energy or particles with its environment, so that the energy of the system does not change with time.

In statistical mechanics, a canonical ensemble is the statistical ensemble that represents the possible states of a mechanical system in thermal equilibrium with a heat bath at a fixed temperature. The system can exchange energy with the heat bath, so that the states of the system will differ in total energy.

<span class="mw-page-title-main">Bekenstein bound</span> Upper limit on entropy in physics

In physics, the Bekenstein bound is an upper limit on the thermodynamic entropy S, or Shannon entropy H, that can be contained within a given finite region of space which has a finite amount of energy—or conversely, the maximal amount of information required to perfectly describe a given physical system down to the quantum level. It implies that the information of a physical system, or the information necessary to perfectly describe that system, must be finite if the region of space and the energy are finite. In computer science this implies that non-finite models such as Turing machines are not realizable as finite devices.

In quantum information theory, quantum discord is a measure of nonclassical correlations between two subsystems of a quantum system. It includes correlations that are due to quantum physical effects but do not necessarily involve quantum entanglement.

The Koopman–von Neumann (KvN) theory is a description of classical mechanics as an operatorial theory similar to quantum mechanics, based on a Hilbert space of complex, square-integrable wavefunctions. As its name suggests, the KvN theory is loosely related to work by Bernard Koopman and John von Neumann in 1931 and 1932, respectively. As explained in this entry, however, the historical origins of the theory and its name are complicated.

In quantum information theory, strong subadditivity of quantum entropy (SSA) is the relation among the von Neumann entropies of various quantum subsystems of a larger quantum system consisting of three subsystems. It is a basic theorem in modern quantum information theory. It was conjectured by D. W. Robinson and D. Ruelle in 1966 and O. E. Lanford III and D. W. Robinson in 1968 and proved in 1973 by E.H. Lieb and M.B. Ruskai, building on results obtained by Lieb in his proof of the Wigner-Yanase-Dyson conjecture.

A quantum heat engine is a device that generates power from the heat flow between hot and cold reservoirs. The operation mechanism of the engine can be described by the laws of quantum mechanics. The first realization of a quantum heat engine was pointed out by Scovil and Schulz-DuBois in 1959, showing the connection of efficiency of the Carnot engine and the 3-level maser. Quantum refrigerators share the structure of quantum heat engines with the purpose of pumping heat from a cold to a hot bath consuming power first suggested by Geusic, Schulz-DuBois, De Grasse and Scovil. When the power is supplied by a laser the process is termed optical pumping or laser cooling, suggested by Wineland and Hänsch. Surprisingly heat engines and refrigerators can operate up to the scale of a single particle thus justifying the need for a quantum theory termed quantum thermodynamics.

Coarse-grained modeling, coarse-grained models, aim at simulating the behaviour of complex systems using their coarse-grained (simplified) representation. Coarse-grained models are widely used for molecular modeling of biomolecules at various granularity levels.

Many-body localization (MBL) is a dynamical phenomenon occurring in isolated many-body quantum systems. It is characterized by the system failing to reach thermal equilibrium, and retaining a memory of its initial condition in local observables for infinite times.

The quantum Fisher information is a central quantity in quantum metrology and is the quantum analogue of the classical Fisher information. The quantum Fisher information of a state with respect to the observable is defined as

References

  1. Deffner, Sebastian and Campbell, Steve. "Quantum Thermodynamics: An introduction to the thermodynamics of quantum information" Morgan & Claypool Publishers (2019), doi.org/10.1088/2053-2571/ab21c6
  2. Binder, F., Correa, L.A., Gogolin, C., Anders, J. and Adesso, G., 2019. Thermodynamics in the Quantum Regime. Fundamental Theories of Physics (Springer, 2018).
  3. Einstein, A. (1905). "Über einen die Erzeugung und Verwandlung des Lichtes betreffenden heuristischen Gesichtspunkt". Annalen der Physik (in German). 322 (6): 132–148. Bibcode:1905AnP...322..132E. doi: 10.1002/andp.19053220607 . ISSN   0003-3804.
  4. John Von Neumann. Mathematical foundations of quantum mechanics. No. 2. Princeton university press, 1955.
  5. 1 2 3 4 Kosloff, Ronnie (2013-05-29). "Quantum Thermodynamics: A Dynamical Viewpoint". Entropy. 15 (12): 2100–2128. arXiv: 1305.2268 . Bibcode:2013Entrp..15.2100K. doi: 10.3390/e15062100 . ISSN   1099-4300.
  6. Lindblad, G. (1976). "On the generators of quantum dynamical semigroups". Communications in Mathematical Physics. 48 (2): 119–130. Bibcode:1976CMaPh..48..119L. doi:10.1007/bf01608499. ISSN   0010-3616. S2CID   55220796.
  7. Gorini, Vittorio (1976). "Completely positive dynamical semigroups of N-level systems". Journal of Mathematical Physics. 17 (5): 821–825. Bibcode:1976JMP....17..821G. doi:10.1063/1.522979. ISSN   0022-2488.
  8. Spohn, H.; Lebowitz, J. Irreversible thermodynamics for quantum systems weakly coupled to thermal reservoirs. Adv. Chem. Phys. 1979, 38, 109.
  9. Alicki, R (1979). "The quantum open system as a model of the heat engine". Journal of Physics A: Mathematical and General. 12 (5): L103–L107. Bibcode:1979JPhA...12L.103A. doi:10.1088/0305-4470/12/5/007. ISSN   0305-4470.
  10. Kosloff, Ronnie (1984-02-15). "A quantum mechanical open system as a model of a heat engine". The Journal of Chemical Physics. 80 (4): 1625–1631. Bibcode:1984JChPh..80.1625K. doi:10.1063/1.446862. ISSN   0021-9606.
  11. Davies, E. B. (1974). "Markovian master equations". Communications in Mathematical Physics. 39 (2): 91–110. Bibcode:1974CMaPh..39...91D. doi:10.1007/bf01608389. ISSN   0010-3616. S2CID   122552267.
  12. Ludovico, María Florencia; Lim, Jong Soo; Moskalets, Michael; Arrachea, Liliana; Sánchez, David (2014-04-21). "Dynamical energy transfer in ac-driven quantum systems". Physical Review B. 89 (16): 161306(R). arXiv: 1311.4945 . Bibcode:2014PhRvB..89p1306L. doi:10.1103/physrevb.89.161306. ISSN   1098-0121. S2CID   119265583.
  13. Esposito, Massimiliano; Ochoa, Maicol A.; Galperin, Michael (2015-02-25). "Quantum Thermodynamics: A Nonequilibrium Green's Function Approach". Physical Review Letters. 114 (8): 080602. arXiv: 1411.1800 . Bibcode:2015PhRvL.114h0602E. doi:10.1103/physrevlett.114.080602. ISSN   0031-9007. PMID   25768745. S2CID   11498686.
  14. Tabakin, Frank (2017-06-03). "Model dynamics for quantum computing". Annals of Physics. 383: 33. arXiv: 1611.00664 . Bibcode:2017AnPhy.383...33T. doi:10.1016/j.aop.2017.04.013. S2CID   119718818.
  15. Beretta, Gian Paolo (2020-05-01). "The fourth law of thermodynamics: steepest entropy ascent". Philosophical Transactions of the Royal Society A. 378 (2170): 20190168. arXiv: 1908.05768 . Bibcode:2020RSPTA.37890168B. doi:10.1098/rsta.2019.0168. ISSN   1471-2962. PMID   32223406. S2CID   201058607.
  16. Lieb, Elliott H.; Yngvason, Jakob (1999). "The physics and mathematics of the second law of thermodynamics". Physics Reports. 310 (1): 1–96. arXiv: cond-mat/9708200 . Bibcode:1999PhR...310....1L. doi:10.1016/s0370-1573(98)00082-9. ISSN   0370-1573. S2CID   119620408.
  17. Gyftopoulos, E. P.; Beretta, G. P. (2005) [1st ed., Macmillan, 1991]. Thermodynamics: Foundations and Applications. Mineola (New York): Dover Publications.
  18. Szilard, L. (1929). "Über die Entropieverminderung in einem thermodynamischen System bei Eingriffen intelligenter Wesen" [On the minimization of entropy in a thermodynamic system with interferences of intelligent beings]. Zeitschrift für Physik (in German). 53 (11–12): 840–856. Bibcode:1929ZPhy...53..840S. doi:10.1007/bf01341281. ISSN   1434-6001. S2CID   122038206.
  19. Brillouin, L. Science and Information Theory; Academic Press: New York, NY, USA, 1956. 107.
  20. Maruyama, Koji; Nori, Franco; Vedral, Vlatko (2009-01-06). "Colloquium: The physics of Maxwell's demon and information". Reviews of Modern Physics. 81 (1): 1–23. arXiv: 0707.3400 . Bibcode:2009RvMP...81....1M. doi:10.1103/revmodphys.81.1. ISSN   0034-6861. S2CID   18436180.
  21. Polkovnikov, Anatoli (2011). "Microscopic diagonal entropy and its connection to basic thermodynamic relations". Annals of Physics. 326 (2): 486–499. arXiv: 0806.2862 . Bibcode:2011AnPhy.326..486P. doi:10.1016/j.aop.2010.08.004. ISSN   0003-4916. S2CID   118412733.
  22. Spohn, H.; Lebowitz, J. Irreversible thermodynamics for quantum systems weakly coupled to thermal reservoirs. Adv. Chem. Phys. 1978, 109, 38.
  23. Levy, Amikam; Kosloff, Ronnie (2014-07-01). "The local approach to quantum transport may violate the second law of thermodynamics". Europhysics Letters. 107 (2): 20004. arXiv: 1402.3825 . Bibcode:2014EL....10720004L. doi:10.1209/0295-5075/107/20004. ISSN   0295-5075. S2CID   118498868.
  24. Gabriele De Chiara, Gabriel Landi, Adam Hewgill, Brendan Reid, Alessandro Ferraro, Augusto J. Roncaglia, and Mauro Antezza, Reconciliation of quantum local master equations with thermodynamics, New Journal of Physics 20, 113024 (2018).
  25. Deng, S.; Chenu, A.; Diao, P.; Li, F.; Yu, S.; Coulamy, I.; del Campo, A; Wu, H. (2018). "Superadiabatic quantum friction suppression in finite-time thermodynamics". Science Advances. 4 (4): eaar5909. arXiv: 1711.00650 . Bibcode:2018SciA....4.5909D. doi:10.1126/sciadv.aar5909. PMC   5922798 . PMID   29719865.
  26. Kosloff, Ronnie; Feldmann, Tova (2002-05-16). "Discrete four-stroke quantum heat engine exploring the origin of friction". Physical Review E. 65 (5): 055102(R). arXiv: physics/0111098 . Bibcode:2002PhRvE..65e5102K. doi:10.1103/physreve.65.055102. ISSN   1063-651X. PMID   12059626. S2CID   9292108.
  27. Plastina, F.; Alecce, A.; Apollaro, T. J. G.; Falcone, G.; Francica, G.; et al. (2014-12-31). "Irreversible Work and Inner Friction in Quantum Thermodynamic Processes". Physical Review Letters. 113 (26): 260601. arXiv: 1407.3441 . Bibcode:2014PhRvL.113z0601P. doi:10.1103/physrevlett.113.260601. ISSN   0031-9007. PMID   25615295. S2CID   9353450.
  28. Landsberg, P. T. (1956-10-01). "Foundations of Thermodynamics". Reviews of Modern Physics. 28 (4): 363–392. Bibcode:1956RvMP...28..363L. doi:10.1103/revmodphys.28.363. ISSN   0034-6861.
  29. Levy, Amikam; Alicki, Robert; Kosloff, Ronnie (2012-06-26). "Quantum refrigerators and the third law of thermodynamics". Physical Review E. 85 (6): 061126. arXiv: 1205.1347 . Bibcode:2012PhRvE..85f1126L. doi:10.1103/physreve.85.061126. ISSN   1539-3755. PMID   23005070. S2CID   24251763.
  30. Bartsch, Christian; Gemmer, Jochen (2009-03-19). "Dynamical Typicality of Quantum Expectation Values". Physical Review Letters. 102 (11): 110403. arXiv: 0902.0927 . Bibcode:2009PhRvL.102k0403B. doi:10.1103/physrevlett.102.110403. ISSN   0031-9007. PMID   19392176. S2CID   34603425.
  31. Goldstein, Sheldon; Lebowitz, Joel L.; Mastrodonato, Christian; Tumulka, Roderich; Zanghì, Nino (2010-05-20). "Normal typicality and von Neumann's quantum ergodic theorem". Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences. 466 (2123): 3203–3224. arXiv: 0907.0108 . Bibcode:2010RSPSA.466.3203G. doi:10.1098/rspa.2009.0635. ISSN   1364-5021. S2CID   816619.
  32. Brandão, Fernando; Horodecki, Michał; Ng, Nelly; Oppenheim, Jonathan; Wehner, Stephanie (2015-02-09). "The second laws of quantum thermodynamics". Proceedings of the National Academy of Sciences. 112 (11): 3275–3279. arXiv: 1305.5278 . Bibcode:2015PNAS..112.3275B. doi: 10.1073/pnas.1411728112 . ISSN   0027-8424. PMC   4372001 . PMID   25675476.
  33. Goold, John; Huber, Marcus; Riera, Arnau; Rio, Lídia del; Skrzypczyk, Paul (2016-02-23). "The role of quantum information in thermodynamics—a topical review". Journal of Physics A: Mathematical and Theoretical. 49 (14): 143001. arXiv: 1505.07835 . Bibcode:2016JPhA...49n3001G. doi: 10.1088/1751-8113/49/14/143001 . ISSN   1751-8113.
  34. M. O. Scully, M. Suhail Zubairy, G. S. Agarwal1, H. Walther, Extracting Work from a Single Heat Bath via Vanishing Quantum Coherence, Science 299, 862-864 (2003).
  35. 1 2 G. Manzano, F. Galve, R. Zambrini, and J. M. R. Parrondo, Entropy production and thermodynamic power of the squeezed thermal reservoir, Phys. Rev. E 93, 052120 (2016).
  36. 1 2 R. J. de Assis, T. M. de Mendonça, C. J. Villas-Boas, A. M. de Souza, R. S. Sarthour, I. S. Oliveira, and N. G. de Almeida, Efficiency of a Quantum Otto Heat Engine Operating under a Reservoir at Effective Negative Temperatures, Phys. Rev. Lett. 122, 240602 (2019).
  37. H. Pothier, S. Guéron, Norman O. Birge, D. Esteve, and M. H. Devoret, Energy Distribution Function of Quasiparticles in Mesoscopic Wires, Phys Rev. Lett. 79, 3490 (1997).
  38. Y.-F. Chen, T. Dirks, G. Al-Zoubi, N. O. Birge, and N. Mason, Nonequilibrium Tunneling Spectroscopy in Carbon Nanotubes, Phys. Rev. Lett. 102, 036804 (2009).
  39. C. Altimiras, H. le Sueur, U. Gennser, A. Cavanna, D. Mailly, and F. Pierre, Tuning Energy Relaxation along Quantum Hall Channels, Phys. Rev. Lett. 105, 226804 (2010).
  40. N. Bronn and N. Maso, Spatial dependence of electron interactions in carbon nanotubes, Phys. Rev. B 88, 161409(R) (2013).
  41. R. Sánchez, J. Splettstoesser, and R. S. Whitney, Nonequilibrium System as a Demon, Phys. Rev. Lett. 123, 216801 (2019).

Further reading