Quantum heat engines and refrigerators

Last updated

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, [1] 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. [2] When the power is supplied by a laser the process is termed optical pumping or laser cooling, suggested by Wineland and Hänsch. [3] [4] [5] 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. [6]

Contents

The 3-level amplifier as a quantum heat engine

The three level amplifier. Levels 1 and 3 are coupled to the hot reservoir. Levels 1 and 2 are coupled to the cold reservoir. Power is obtained when there is population inversion between levels 3 and 2. Three-level-amp.pdf
The three level amplifier. Levels 1 and 3 are coupled to the hot reservoir. Levels 1 and 2 are coupled to the cold reservoir. Power is obtained when there is population inversion between levels 3 and 2.

The three-level-amplifier is the template of a quantum device. It operates by employing a hot and cold bath to maintain population inversion between two energy levels which is used to amplify light by stimulated emission [7] The ground state level (1-g) and the excited level (3-h) are coupled to a hot bath of temperature . The energy gap is . When the population on the levels equilibrate

where is the Planck constant and is the Boltzmann constant. The cold bath of temperature couples the ground (1-g) to an intermediate level (2-c) with energy gap . When levels 2-c and 1-g equilibrate then

.

The device operates as an amplifier when levels (3-h) and (2-c) are coupled to an external field of frequency . For optimal resonance conditions . The efficiency of the amplifier in converting heat to power is the ratio of work output to heat input:

.

Amplification of the field is possible only for positive gain (population inversion) . This is equivalent to . Inserting this expression into the efficiency formula leads to:

where is the Carnot cycle efficiency. Equality is obtained under a zero gain condition . The relation between the quantum amplifier and the Carnot efficiency was first pointed out by Scovil and Schultz-DuBois: [1]

Reversing the operation driving heat from the cold bath to the hot bath by consuming power constitutes a refrigerator. The efficiency of the refrigerator defined as the coefficient of performance (COP) for the reversed device is:

Types

Quantum devices can operate either continuously or by a reciprocating cycle. Continuous devices include solar cells converting solar radiation to electrical power, thermoelectric where the output is current and lasers where the output power is coherent light. The primary example of a continuous refrigerator is optical pumping and laser cooling. [8] [9] Similarly to classical reciprocating engines, quantum heat engines also have a cycle that is divided into different strokes. A stroke is time segment in which a certain operation takes place (e.g. thermalization, or work extraction). Two adjacent strokes do not commute with each other. The most common reciprocating heat machines are the four-stroke machine, and the two-stroke machine. Reciprocating devices have been suggested operating either by the Carnot cycle [10] [11] or the Otto cycle. [12]

In both types the quantum description allows to obtain equation of motion for the working medium and the heat flow from the reservoirs.

Quantum reciprocating heat engine and refrigerator

Quantum versions of most of the common thermodynamic cycles have been studied, for example the Carnot cycle, [10] [11] [13] Stirling cycle [14] and Otto cycle. [12] [15]

The Otto cycle can serve as a template for other reciprocating cycles.

Quantum Otto cycle shown in the Entropy
O
{\displaystyle \Omega }
plane where the energy entropy and the Von Neumann entropy are displayed.
O
{\displaystyle \Omega }
is the internal frequency of the device and is controlled externally. It mimics the inverse volume in the Otto cycle. The red and blue lines are the hot and cold isochores. The cycle represents a heat pump. Q-otto-cycle.pdf
Quantum Otto cycle shown in the Entropy plane where the energy entropy and the Von Neumann entropy are displayed. is the internal frequency of the device and is controlled externally. It mimics the inverse volume in the Otto cycle. The red and blue lines are the hot and cold isochores. The cycle represents a heat pump.

It is composed of the following four segments:

The propagator of the four stroke cycle becomes , which is the ordered product of the segment propagators:

The propagators are linear operators defined on a vector space which completely determines the state of the working medium. Common to all thermodynamic cycles the consecutive segment propagators do not commute . Commuting propagators will lead to zero power.

In a reciprocating quantum heat engine the working medium is a quantum system such as spin systems [16] or an harmonic oscillator. [17] For maximum power the cycle time should be optimized. There are two basic timescales in the reciprocating refrigerator the cycle time and the internal timescale . In general when the engine operates in quasi-adiabatic conditions. The only quantum effect can be found at low temperatures where the unit of energy of the device becomes instead of . The efficiency at this limit is , always smaller than the Carnot efficiency . At high temperature and for the harmonic working medium the efficiency at maximum power becomes which is the endoreversible thermodynamics result. [17]

For shorter cycle times the working medium cannot follow adiabatically the change in the external parameter. This leads to friction-like phenomena. Extra power is required to drive the system faster. The signature of such dynamics is the development of coherence causing extra dissipation. Surprisingly the dynamics leading to friction is quantized meaning that frictionless solutions to the adiabatic expansion/compression can be found in finite time. [18] [19] As a result, optimization has to be carried out only with respect to the time allocated to heat transport. In this regime the quantum feature of coherence degrades the performance. Optimal frictionless performance is obtained when the coherence can be cancelled.

The shortest cycle times , sometimes termed sudden cycles, [20] have universal features. In this case coherence contributes to the cycles power.

A two-stroke engine quantum cycle equivalent to the Otto cycle based on two qubits has been proposed. The first qubit has frequency and the second . The cycle is composed of a first stroke of partial equilibration of the two qubits with the hot and cold bath in parallel. The second power stroke is composed of a partial or full swap between the qubits. The swap operation is generated by a unitary transformation which preserves the entropy as a result it is a pure power stroke. [21] [22]

The quantum Otto cycle refrigerators shares the same cycle with magnetic refrigeration. [23]

Continuous quantum engines

Continuous quantum engines are the quantum analogues of turbines. The work output mechanism is coupling to an external periodic field, typically the electromagnetic field. Thus the heat engine is a model for a laser. [9] The models differ by the choice of their working substance and heat source and sink. Externally driven two-level, [24] three level [25] four-level [26] [27] and coupled harmonic oscillators [28] have been studied.

The periodic driving splits the energy level structure of the working medium. This splitting allows the two level engine to couple selectively to the hot and cold baths and produce power. On the other hand, ignoring this splitting in the derivation of the equation of motion will violate the second law of thermodynamics. [29]

Non thermal fuels have been considered for quantum heat engines. The idea is to increase the energy content of the hot bath without increasing its entropy. This can be achieved by employing coherence [30] or a squeezed thermal bath. [31] These devices do not violate the second law of thermodynamics.

Equivalence of reciprocating and continuous heat machines in the quantum regime

Two-stroke, Four-stroke, and continuous machine are very different from each other. However it was shown [32] that there is a quantum regime where all these machines become thermodynamically equivalent to each other. While the intra cycle dynamics in the equivalence regime is very different in different engine types, when the cycle is completed they all turn out to provide the same amount of work and consume the same amount of heat (hence they share the same efficiency as well). This equivalence is associated with a coherent work extraction mechanism and has no classical analogue. These quantum features have been demonstrated experimentally. [33]

Heat engines and open quantum systems

The elementary example operates under quasi equilibrium conditions. Its main quantum feature is the discrete energy level structure. More realistic devices operate out of equilibrium possessing friction heat leaks and finite heat flow. Quantum thermodynamics supplies a dynamical theory required for systems out of equilibrium such as heat engines, thus, inserting dynamics into thermodynamics. The theory of open quantum systems constitutes the basic theory. For heat engines a reduced description of the dynamics of the working substance is sought, tracing out the hot and cold baths. The starting point is the general Hamiltonian of the combined systems:

and the system Hamiltonian is time dependent. A reduced description leads to the equation of motion of the system:

where is the density operator describing the state of the working medium and is the generator of dissipative dynamics which includes the heat transport terms from the baths. Using this construction, the total change in energy of the sub-system becomes:

leading to the dynamical version of the first law of thermodynamics: [6]

The rate of entropy production becomes:

The global structure of quantum mechanics is reflected in the derivation of the reduced description. A derivation which is consistent with the laws of thermodynamics is based on the weak coupling limit. A thermodynamical idealization assumes that the system and the baths are uncorrelated, meaning that the total state of the combined system becomes a tensor product at all times:

Under these conditions the dynamical equations of motion become: where is the Liouville superoperator described in terms of the system's Hilbert space, where the reservoirs are described implicitly. Within the formalism of quantum open system, can take the form of the Gorini-Kossakowski-Sudarshan-Lindblad (GKS-L) Markovian generator or also known just as Lindblad equation . [34] Theories beyond the weak coupling regime have been proposed. [35] [36] [37]

The quantum absorption refrigerator

The absorption refrigerator is of unique importance in setting an autonomous quantum device. Such a device requires no external power and operates without external intervention in scheduling the operations . [38] [39] [40] The basic construct includes three baths; a power bath, a hot bath and a cold bath. The tricycle model is the template for the absorption refrigerator.

Quantum tricycle absorption refrigerator. The device is composed from three baths where
T
d
>=
T
h
>=
T
c
{\displaystyle T_{\text{d}}\geq T_{\text{h}}\geq T_{\text{c}}}
. Heat flows from the power reservoir and cold bath to the hot bath. Tricyle-1.png
Quantum tricycle absorption refrigerator. The device is composed from three baths where . Heat flows from the power reservoir and cold bath to the hot bath.

The tricycle engine has a generic structure. The basic model consists of three thermal baths: A hot bath with temperature , a cold bath with temperature and a work bath with temperature .

Each bath is connected to the engine via a frequency filter which can be modeled by three oscillators:

where , and are the filter frequencies on resonance .

The device operates as a refrigerator by removing an excitation from the cold bath as well as from the work bath and generating an excitation in the hot bath. The term in the Hamiltonian is non linear and crucial for an engine or a refrigerator.

where is the coupling strength.

The first-law of thermodynamics represents the energy balance of heat currents originating from the three baths and collimating on the system:

At steady state no heat is accumulated in the tricycle, thus . In addition, in steady state the entropy is only generated in the baths, leading to the second law of thermodynamics:

This version of the second-law is a generalisation of the statement of Clausius theorem; heat does not flow spontaneously from cold to hot bodies. When the temperature , no entropy is generated in the power bath. An energy current with no accompanying entropy production is equivalent to generating pure power: , where is the power output.

Quantum refrigerators and 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: [41]

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; [42]

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 introduce 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.

Related Research Articles

<span class="mw-page-title-main">Uncertainty principle</span> Foundational principle in quantum physics

The uncertainty principle, also known as Heisenberg's indeterminacy principle, is a fundamental concept in quantum mechanics. It states that there is a limit to the precision with which certain pairs of physical properties, such as position and momentum, can be simultaneously known. In other words, the more accurately one property is measured, the less accurately the other property can be known.

The quantum Hall effect is a quantized version of the Hall effect which is observed in two-dimensional electron systems subjected to low temperatures and strong magnetic fields, in which the Hall resistance Rxy exhibits steps that take on the quantized values

<span class="mw-page-title-main">Zero-point energy</span> Lowest possible energy of a quantum system or field

Zero-point energy (ZPE) is the lowest possible energy that a quantum mechanical system may have. Unlike in classical mechanics, quantum systems constantly fluctuate in their lowest energy state as described by the Heisenberg uncertainty principle. Therefore, even at absolute zero, atoms and molecules retain some vibrational motion. Apart from atoms and molecules, the empty space of the vacuum also has these properties. According to quantum field theory, the universe can be thought of not as isolated particles but continuous fluctuating fields: matter fields, whose quanta are fermions, and force fields, whose quanta are bosons. All these fields have zero-point energy. These fluctuating zero-point fields lead to a kind of reintroduction of an aether in physics since some systems can detect the existence of this energy. However, this aether cannot be thought of as a physical medium if it is to be Lorentz invariant such that there is no contradiction with Einstein's theory of special relativity.

In physics, specifically in quantum mechanics, a coherent state is the specific quantum state of the quantum harmonic oscillator, often described as a state that has dynamics most closely resembling the oscillatory behavior of a classical harmonic oscillator. It was the first example of quantum dynamics when Erwin Schrödinger derived it in 1926, while searching for solutions of the Schrödinger equation that satisfy the correspondence principle. The quantum harmonic oscillator arise in the quantum theory of a wide range of physical systems. For instance, a coherent state describes the oscillating motion of a particle confined in a quadratic potential well. The coherent state describes a state in a system for which the ground-state wavepacket is displaced from the origin of the system. This state can be related to classical solutions by a particle oscillating with an amplitude equivalent to the displacement.

<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.

The fluctuation–dissipation theorem (FDT) or fluctuation–dissipation relation (FDR) is a powerful tool in statistical physics for predicting the behavior of systems that obey detailed balance. Given that a system obeys detailed balance, the theorem is a proof that thermodynamic fluctuations in a physical variable predict the response quantified by the admittance or impedance of the same physical variable, and vice versa. The fluctuation–dissipation theorem applies both to classical and quantum mechanical systems.

<span class="mw-page-title-main">Polaron</span> Quasiparticle in condensed matter physics

A polaron is a quasiparticle used in condensed matter physics to understand the interactions between electrons and atoms in a solid material. The polaron concept was proposed by Lev Landau in 1933 and Solomon Pekar in 1946 to describe an electron moving in a dielectric crystal where the atoms displace from their equilibrium positions to effectively screen the charge of an electron, known as a phonon cloud. This lowers the electron mobility and increases the electron's effective mass.

<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.

<span class="mw-page-title-main">Jaynes–Cummings model</span> Model in quantum optics

The Jaynes–Cummings model is a theoretical model in quantum optics. It describes the system of a two-level atom interacting with a quantized mode of an optical cavity, with or without the presence of light. It was originally developed to study the interaction of atoms with the quantized electromagnetic field in order to investigate the phenomena of spontaneous emission and absorption of photons in a cavity.

In spectroscopy, the Autler–Townes effect, is a dynamical Stark effect corresponding to the case when an oscillating electric field is tuned in resonance to the transition frequency of a given spectral line, and resulting in a change of the shape of the absorption/emission spectra of that spectral line. The AC Stark effect was discovered in 1955 by American physicists Stanley Autler and Charles Townes.

The Gross–Pitaevskii equation describes the ground state of a quantum system of identical bosons using the Hartree–Fock approximation and the pseudopotential interaction model.

Circuit quantum electrodynamics provides a means of studying the fundamental interaction between light and matter. As in the field of cavity quantum electrodynamics, a single photon within a single mode cavity coherently couples to a quantum object (atom). In contrast to cavity QED, the photon is stored in a one-dimensional on-chip resonator and the quantum object is no natural atom but an artificial one. These artificial atoms usually are mesoscopic devices which exhibit an atom-like energy spectrum. The field of circuit QED is a prominent example for quantum information processing and a promising candidate for future quantum computation.

The phase-space formulation of quantum mechanics places the position and momentum variables on equal footing in phase space. In contrast, the Schrödinger picture uses the position or momentum representations. The two key features of the phase-space formulation are that the quantum state is described by a quasiprobability distribution and operator multiplication is replaced by a star product.

<span class="mw-page-title-main">Superradiant phase transition</span> Process in quantum optics

In quantum optics, a superradiant phase transition is a phase transition that occurs in a collection of fluorescent emitters, between a state containing few electromagnetic excitations and a superradiant state with many electromagnetic excitations trapped inside the emitters. The superradiant state is made thermodynamically favorable by having strong, coherent interactions between the emitters.

<span class="mw-page-title-main">Quantum thermodynamics</span> Study of the relations between thermodynamics and quantum mechanics

Quantum thermodynamics 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 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. 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.

Electric dipole spin resonance (EDSR) is a method to control the magnetic moments inside a material using quantum mechanical effects like the spin–orbit interaction. Mainly, EDSR allows to flip the orientation of the magnetic moments through the use of electromagnetic radiation at resonant frequencies. EDSR was first proposed by Emmanuel Rashba.

In quantum probability, the Belavkin equation, also known as Belavkin-Schrödinger equation, quantum filtering equation, stochastic master equation, is a quantum stochastic differential equation describing the dynamics of a quantum system undergoing observation in continuous time. It was derived and henceforth studied by Viacheslav Belavkin in 1988.

<span class="mw-page-title-main">Cavity optomechanics</span>

Cavity optomechanics is a branch of physics which focuses on the interaction between light and mechanical objects on low-energy scales. It is a cross field of optics, quantum optics, solid-state physics and materials science. The motivation for research on cavity optomechanics comes from fundamental effects of quantum theory and gravity, as well as technological applications.

The hierarchical equations of motion (HEOM) technique derived by Yoshitaka Tanimura and Ryogo Kubo in 1989, is a non-perturbative approach developed to study the evolution of a density matrix of quantum dissipative systems. The method can treat system-bath interaction non-perturbatively as well as non-Markovian noise correlation times without the hindrance of the typical assumptions that conventional Redfield (master) equations suffer from such as the Born, Markovian and rotating-wave approximations. HEOM is applicable even at low temperatures where quantum effects are not negligible.

The continuous spontaneous localization (CSL) model is a spontaneous collapse model in quantum mechanics, proposed in 1989 by Philip Pearle. and finalized in 1990 Gian Carlo Ghirardi, Philip Pearle and Alberto Rimini.

References

  1. 1 2 Scovil, H. E. D.; Schulz-DuBois, E. O. (1959). "Three-Level Masers as Heat Engines". Physical Review Letters. 2 (6): 262–263. Bibcode:1959PhRvL...2..262S. doi:10.1103/PhysRevLett.2.262. ISSN   0031-9007.
  2. Geusic, J. E.; Bois, E. O. Schulz-Du; De Grasse, R. W.; Scovil, H. E. D. (1959). "Three Level Spin Refrigeration and Maser Action at 1500 mc/sec". Journal of Applied Physics. 30 (7): 1113–1114. Bibcode:1959JAP....30.1113G. doi:10.1063/1.1776991. ISSN   0021-8979.
  3. D. J. Wineland and H. Dehmelt, Bull. Am. Phys. Soc. 20, 637 (1975)
  4. Hänsch, T.W.; Schawlow, A.L. (1975). "Cooling of gases by laser radiation". Optics Communications. 13 (1): 68–69. Bibcode:1975OptCo..13...68H. doi: 10.1016/0030-4018(75)90159-5 . ISSN   0030-4018.
  5. Letokhov, V.S.; Minogin, V.G.; Pavlik, B.D. (1976). "Cooling and trapping of atoms and molecules by a resonant laser field". Optics Communications. 19 (1): 72–75. Bibcode:1976OptCo..19...72L. doi:10.1016/0030-4018(76)90388-6. ISSN   0030-4018.
  6. 1 2 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.
  7. Yariv, Amnon (1989). Quantum Electronics, 3rd ed., Wiley. ISBN   0-471-60997-8
  8. Narevicius, Edvardas; Bannerman, S Travis; Raizen, Mark G (2009). "Single-photon molecular cooling". New Journal of Physics. 11 (5): 055046. arXiv: 0808.1383 . Bibcode:2009NJPh...11e5046N. doi: 10.1088/1367-2630/11/5/055046 . ISSN   1367-2630.
  9. 1 2 Kosloff, Ronnie; Levy, Amikam (2014). "Quantum Heat Engines and Refrigerators: Continuous Devices". Annual Review of Physical Chemistry. 65 (1): 365–393. arXiv: 1310.0683 . Bibcode:2014ARPC...65..365K. doi:10.1146/annurev-physchem-040513-103724. ISSN   0066-426X. PMID   24689798. S2CID   25266545.
  10. 1 2 Geva, Eitan; Kosloff, Ronnie (1992). "A quantum-mechanical heat engine operating in finite time. A model consisting of spin-1/2 systems as the working fluid". The Journal of Chemical Physics. 96 (4): 3054–3067. Bibcode:1992JChPh..96.3054G. doi:10.1063/1.461951. ISSN   0021-9606.
  11. 1 2 Bender, Carl M; Brody, Dorje C; Meister, Bernhard K (2000). "Quantum mechanical Carnot engine". Journal of Physics A: Mathematical and General. 33 (24): 4427–4436. arXiv: quant-ph/0007002 . Bibcode:2000JPhA...33.4427B. doi:10.1088/0305-4470/33/24/302. ISSN   0305-4470. S2CID   5335.
  12. 1 2 Feldmann, Tova; Kosloff, Ronnie (2000). "Performance of discrete heat engines and heat pumps in finite time". Physical Review E. 61 (5): 4774–4790. arXiv: physics/0003007 . Bibcode:2000PhRvE..61.4774F. doi:10.1103/PhysRevE.61.4774. ISSN   1063-651X. PMID   11031518. S2CID   2277942.
  13. Quan, H. T.; Liu, Yu-xi; Sun, C. P.; Nori, Franco (2007). "Quantum thermodynamic cycles and quantum heat engines". Physical Review E. 76 (3): 031105. arXiv: quant-ph/0611275 . Bibcode:2007PhRvE..76c1105Q. doi:10.1103/PhysRevE.76.031105. ISSN   1539-3755. PMID   17930197. S2CID   3009953.
  14. Wu, F.; Chen, L.; Sun, F.; Wu, C.; Zhu, Yonghong (1998). "Performance and optimization criteria for forward and reverse quantum Stirling cycles". Energy Conversion and Management. 39 (8): 733–739. Bibcode:1998ECM....39..733W. doi:10.1016/S0196-8904(97)10037-1. ISSN   0196-8904.
  15. Kieu, T. D. (2006). "Quantum heat engines, the second law and Maxwell's daemon". The European Physical Journal D. 39 (1): 115–128. arXiv: quant-ph/0311157 . Bibcode:2006EPJD...39..115K. doi:10.1140/epjd/e2006-00075-5. ISSN   1434-6060. S2CID   119382163.
  16. Feldmann, Tova; Kosloff, Ronnie (2003). "Quantum four-stroke heat engine: Thermodynamic observables in a model with intrinsic friction". Physical Review E. 68 (1): 016101. arXiv: quant-ph/0303046 . Bibcode:2003PhRvE..68a6101F. doi:10.1103/PhysRevE.68.016101. ISSN   1063-651X. PMID   12935194. S2CID   23777311.
  17. 1 2 Rezek, Yair; Kosloff, Ronnie (2006). "Irreversible performance of a quantum harmonic heat engine". New Journal of Physics. 8 (5): 83. arXiv: quant-ph/0601006 . Bibcode:2006NJPh....8...83R. doi: 10.1088/1367-2630/8/5/083 . ISSN   1367-2630.
  18. Campo, A. del; Goold, J.; Paternostro, M. (2014). "More bang for your buck: Super-adiabatic quantum engines". Scientific Reports. 4 (1): 6208. Bibcode:2014NatSR...4E6208C. doi: 10.1038/srep06208 . ISSN   2045-2322. PMC   4147366 . PMID   25163421.
  19. Beau, Mathieu; Jaramillo, Juan; del Campo, Adolfo (2016). "Scaling-Up Quantum Heat Engines Efficiently via Shortcuts to Adiabaticity". Entropy. 18 (5): 168. arXiv: 1603.06019 . Bibcode:2016Entrp..18..168B. doi: 10.3390/e18050168 . ISSN   1099-4300.
  20. Feldmann, Tova; Kosloff, Ronnie (2012). "Short time cycles of purely quantum refrigerators". Physical Review E. 85 (5): 051114. arXiv: 1204.4059 . Bibcode:2012PhRvE..85e1114F. doi:10.1103/PhysRevE.85.051114. ISSN   1539-3755. PMID   23004710. S2CID   31174895.
  21. Allahverdyan, Armen E.; Hovhannisyan, Karen; Mahler, Guenter (2010). "Optimal refrigerator". Physical Review E. 81 (5): 051129. arXiv: 1007.4307 . Bibcode:2010PhRvE..81e1129A. doi:10.1103/PhysRevE.81.051129. ISSN   1539-3755. PMID   20866207. S2CID   12750223.
  22. Uzdin, Raam; Kosloff, Ronnie (2014). "The multilevel four-stroke swap engine and its environment". New Journal of Physics. 16 (9): 095003. arXiv: 1404.6182 . Bibcode:2014NJPh...16i5003U. doi: 10.1088/1367-2630/16/9/095003 . ISSN   1367-2630.
  23. Shirron, Peter J.; McCammon, Dan (2014). "Salt pill design and fabrication for adiabatic demagnetization refrigerators". Cryogenics. 62: 163–171. Bibcode:2014Cryo...62..163S. doi:10.1016/j.cryogenics.2014.03.022. ISSN   0011-2275.
  24. Gelbwaser-Klimovsky, D.; Alicki, R.; Kurizki, G. (2013). "Minimal universal quantum heat machine". Physical Review E. 87 (1): 012140. arXiv: 1209.1190 . Bibcode:2013PhRvE..87a2140G. doi:10.1103/PhysRevE.87.012140. ISSN   1539-3755. PMID   23410316. S2CID   18826566.
  25. Geva, Eitan; Kosloff, Ronnie (1996). "The quantum heat engine and heat pump: An irreversible thermodynamic analysis of the three-level amplifier". The Journal of Chemical Physics. 104 (19): 7681–7699. Bibcode:1996JChPh.104.7681G. doi:10.1063/1.471453. ISSN   0021-9606.
  26. Scully, M. O.; Chapin, K. R.; Dorfman, K. E.; Kim, M. B.; Svidzinsky, A. (2011). "Quantum heat engine power can be increased by noise-induced coherence". Proceedings of the National Academy of Sciences. 108 (37): 15097–15100. Bibcode:2011PNAS..10815097S. doi: 10.1073/pnas.1110234108 . ISSN   0027-8424. PMC   3174605 . PMID   21876187.
  27. Harbola, Upendra; Rahav, Saar; Mukamel, Shaul (2012). "Quantum heat engines: A thermodynamic analysis of power and efficiency". EPL (Europhysics Letters). 99 (5): 50005. Bibcode:2012EL.....9950005H. doi:10.1209/0295-5075/99/50005. ISSN   0295-5075. S2CID   13833767.
  28. Kosloff, Ronnie (1984). "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.
  29. Szczygielski, Krzysztof; Gelbwaser-Klimovsky, David; Alicki, Robert (2013). "Markovian master equation and thermodynamics of a two-level system in a strong laser field". Physical Review E. 87 (1): 012120. arXiv: 1211.5665 . Bibcode:2013PhRvE..87a2120S. doi:10.1103/PhysRevE.87.012120. ISSN   1539-3755. PMID   23410296. S2CID   25511420.
  30. Scully, M. O. (2003). "Extracting Work from a Single Heat Bath via Vanishing Quantum Coherence". Science. 299 (5608): 862–864. Bibcode:2003Sci...299..862S. doi: 10.1126/science.1078955 . ISSN   0036-8075. PMID   12511655. S2CID   120884236.
  31. Roßnagel, J.; Abah, O.; Schmidt-Kaler, F.; Singer, K.; Lutz, E. (2014). "Nanoscale Heat Engine Beyond the Carnot Limit". Physical Review Letters. 112 (3): 030602. arXiv: 1308.5935 . Bibcode:2014PhRvL.112c0602R. doi:10.1103/PhysRevLett.112.030602. ISSN   0031-9007. PMID   24484127. S2CID   1826585.
  32. Uzdin, Raam; Levy, Amikam; Kosloff, Ronnie (2015). "Equivalence of Quantum Heat Machines, and Quantum-Thermodynamic Signatures". Physical Review X. 5 (3): 031044. arXiv: 1502.06592 . Bibcode:2015PhRvX...5c1044U. doi: 10.1103/PhysRevX.5.031044 . ISSN   2160-3308.
  33. Klatzow, James; Becker, Jonas N; Ledingham, Patrick M; Weinzetl, Christian; Kaczmarek, Krzysztof T; Saunders, Dylan J; Nunn, Joshua; Walmsley, Ian A; Uzdin, Raam; Poem, Eilon (2019). "Experimental demonstration of quantum effects in the operation of microscopic heat engines". Physical Review Letters. 122 (11): 110601. arXiv: 1710.08716 . Bibcode:2019PhRvL.122k0601K. doi:10.1103/PhysRevLett.122.110601. ISSN   1079-7114. PMID   30951320. S2CID   206318713.
  34. Kosloff, Ronnie (2013). "Quantum Thermodynamics: A Dynamical Viewpoint". Entropy. 15 (12): 2100–2128. arXiv: 1305.2268 . Bibcode:2013Entrp..15.2100K. doi: 10.3390/e15062100 . ISSN   1099-4300.
  35. Ludovico, M. F.; Lim, J. S.; Moskalets, M.; Arrachea, L.; Sánchez, D. (21 April 2014). "Dynamical energy transfer in ac-driven quantum systems". Physical Review B. 89 (16): 161306. Bibcode:2014PhRvB..89p1306L. doi:10.1103/PhysRevB.89.161306. hdl: 10261/116187 . S2CID   119265583.
  36. Gallego, R; Riera, A; Eisert, J (2014). "Thermal machines beyond the weak coupling regime". New Journal of Physics. 16 (12): 125009. arXiv: 1310.8349 . Bibcode:2014NJPh...16l5009G. doi: 10.1088/1367-2630/16/12/125009 . ISSN   1367-2630.
  37. Esposito, Massimiliano; Ochoa, Maicol A.; Galperin, Michael (2015). "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.
  38. Palao, José P.; Kosloff, Ronnie; Gordon, Jeffrey M. (2001). "Quantum thermodynamic cooling cycle". Physical Review E. 64 (5): 056130. arXiv: quant-ph/0106048 . Bibcode:2001PhRvE..64e6130P. doi:10.1103/PhysRevE.64.056130. ISSN   1063-651X. PMID   11736037. S2CID   8201978.
  39. Linden, Noah; Popescu, Sandu; Skrzypczyk, Paul (2010). "How Small Can Thermal Machines Be? The Smallest Possible Refrigerator". Physical Review Letters. 105 (13): 130401. arXiv: 0908.2076 . Bibcode:2010PhRvL.105m0401L. doi:10.1103/PhysRevLett.105.130401. ISSN   0031-9007. PMID   21230755. S2CID   2707740.
  40. Levy, Amikam; Kosloff, Ronnie (2012). "Quantum Absorption Refrigerator". Physical Review Letters. 108 (7): 070604. arXiv: 1109.0728 . Bibcode:2012PhRvL.108g0604L. doi:10.1103/PhysRevLett.108.070604. ISSN   0031-9007. PMID   22401189. S2CID   6981288.
  41. Landsberg, P. T. (1956). "Foundations of Thermodynamics". Reviews of Modern Physics. 28 (4): 363–392. Bibcode:1956RvMP...28..363L. doi:10.1103/RevModPhys.28.363. ISSN   0034-6861.
  42. Levy, Amikam; Alicki, Robert; Kosloff, Ronnie (2012). "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.

Further reading

Deffner, Sebastian and Campbell, Steve. "Quantum Thermodynamics: An introduction to the thermodynamics of quantum information", (Morgan & Claypool Publishers, 2019). [1]

F. Binder, L. A. Correa, C. Gogolin, J. Anders, G. Adesso (eds.) "Thermodynamics in the Quantum Regime. Fundamental Aspects and New Directions." (Springer 2018)

Gemmer, Jochen, M. Michel, and Günter Mahler. "Quantum thermodynamics. Emergence of thermodynamic behavior within composite quantum systems. 2." (2009).

Petruccione, Francesco, and Heinz-Peter Breuer. The theory of open quantum systems. Oxford university press, 2002.

  1. Deffner, Sebastian (2019). Quantum Thermodynamics. doi:10.1088/2053-2571/ab21c6. ISBN   978-1-64327-658-8. S2CID   195791624.