Research aimed at elucidating the foundations of quantum theory can have a direct impact on quantum technology. Two examples illustrate this potential: (1) the coupling of quantum systems to arbitrary classical environments that can be described by irreversible thermodynamics. In the spirit of Dirac’s replacement of classical Poisson brackets by commutators, a thermodynamically consistent coupling of quantum and classical systems can be obtained by quantization of the geometric structure of classical irreversible thermodynamics. (2) The stochastic bra-ket interpretation of quantum mechanics, which is obtained by unraveling density matrices in terms of bra-ket pairs of stochastic jump processes in Hilbert space. It offers an alternative realization of entanglement and avoids paradoxes by imposing severe but natural restrictions on the types of systems to which quantum mechanics can be applied.

Any practical quantum device must interact with the classical world of our direct experience in some way. For instance, a quantum sensor needs some kind of display, and a quantum computer should presumably interface with a classical computer to achieve its full potential. A Hamiltonian coupling between quantum systems and their classical environment is expected to be the most effective method for transferring “results” gained by quantum sensors or computers. However, a detailed understanding of dissipative couplings between quantum and classical systems is also important, if only to minimize undesirable dissipative interactions.

The good news of the first part of this paper is that there exists a general thermodynamic framework for coupling quantum systems to their classical environments, which are assumed to evolve according to the laws of reversible and irreversible dynamics. This framework has been established by quantizing a geometric formulation of classical irreversible thermodynamics, offering a significant extension of the theory of open quantum systems.1,2

The development of new and improved quantum devices would undoubtedly benefit from a more intuitive understanding of quantum mechanics. In particular, entanglement, which embodies the holistic nature of quantum mechanics, is counterintuitive but crucial. In the standard approach, it leads to several well-known paradoxes. When a quantum system is divided into two subsystems, such as two particles or groups of particles, entanglement is the phenomenon that each subsystem cannot be described independently of the state of the other subsystem, even when the subsystems are separated by a large distance.

After 100 years of quantum mechanics, rather than adhering to dogmatism and relying solely on its mathematical framework, we should demand a convincing interpretation. Should we not consider the so-called measurement problem of quantum mechanics a potential obstacle for the advancement of quantum technology?3–6 

At least Feynman was amusingly irritated when he began contemplating what quantum computers could be good for,7 “we always have had a great deal of difficulty in understanding the worldview that quantum mechanics represents. At least I do, because I am an old enough man that I haven't got to the point that this stuff is obvious to me. Okay, I still get nervous with it … you know how it always is. every new idea, it takes a generation or two until it becomes obvious that there's no real problem. It has not yet become obvious to me that there's no real problem. I cannot define the real problem, therefore, I suspect there's no real problem, but I'm not sure there's no real problem.”

The good news in the second part of this paper is that there exists a new interpretation of quantum mechanics, which offers a fresh perspective on entanglement and eliminates the paradoxes that arise in the standard approach to quantum mechanics.

The emergence of irreversibility has been intensely investigated and heavily debated for ∼150 years. Boltzmann’s transport equation for rarefied gases8 marks the first milestone in this development. It is an irreversible equation for the single-particle probability density in position and momentum space based on the collision laws obtained from the Hamiltonian dynamics of gas particles. It also implies an evolution equation for entropy. By the end of the 19th century, Boltzmann had clearly understood the probabilistic nature of the second law of thermodynamics, recognizing that violations would never be observed for macroscopic systems but could become noticeable in very small systems.

Fluctuation–dissipation relations9–13 and projection-operator methods14–18 played a significant role in advancing the field of irreversible thermodynamics. In the 1960s, the foundational principles for formulating linear irreversible thermodynamics were well-established and documented in a classical textbook.19 

An elegant geometric formulation of classical nonequilibrium thermodynamics, initiated by Grmela20,21 in 1984, has led to the so-called GENERIC framework (general equation for the nonequilibrium reversible–irreversible coupling):22,23 Reversible dynamics is generated by energy via a Poisson bracket, whereas irreversible dynamics is generated by entropy via a dissipative bracket.

Understanding the emergence of irreversibility on the quantum level is expected to be considerably more challenging, certainly not easier than for classical systems. We aim to avoid limitations imposed by perturbation theory, simplistic models like reservoirs composed of harmonic oscillators, or approximations of unclear validity.

Lindblad formulated a quantum master equation for the density matrix of a dissipative quantum system based on the assumptions of linearity and complete positivity.24 Grabert has used the projection-operator method to derive a quantum master equation that is nonlinear in the density matrix.18,25 Both equations address the dissipative coupling of a quantum system to a bath. In many applications, the classical environment should not be limited to a heat bath. Ideally, the coupling of a quantum system to an arbitrary classical nonequilibrium system would be desirable.

A systematic framework for quantum systems in contact with finite quantum heat reservoirs has been established in a pioneering paper.26 This approach elaborates the meaning of entropy production and sheds light on the emergence of irreversibility in the limit of large heat reservoirs. More recent developments are summarized in a broad collection of over 40 papers27 and in a recent review article,28 both of which emphasize fluctuation theorems and information-theoretic aspects while also addressing experimental achievements and practical applications.

The geometric structure of the GENERIC framework offers the opportunity to extend Dirac’s approach to quantization from Hamiltonian to dissipative systems. Instead of deriving suitable master equations for dissipative quantum systems emerging from the reversible equations of quantum mechanics, a quantization procedure, in the spirit of replacing Poisson brackets by commutators, is applied to dissipative classical systems. As illustrated in Fig. 1, this idea eliminates the most challenging task of explaining the emergence of irreversibility at the quantum level.

FIG. 1.

Dirac-style quantization of dissipative classical systems for avoiding the challenging task of explaining the emergence of irreversibility at the quantum level.

FIG. 1.

Dirac-style quantization of dissipative classical systems for avoiding the challenging task of explaining the emergence of irreversibility at the quantum level.

Close modal

As the von Neumann entropy is readily available as a generator of irreversible dynamics for quantum systems described by density matrices, one only needs to find a quantization rule for the dissipative bracket, analogous to Dirac’s replacement of Poisson brackets with commutators.

This idea has been pursued in Ref. 29 and further formalized and generalized in Ref. 30. The argumentation and notation in those papers are very abstract because the emphasis is on the deep structural features of the procedure. Here, we offer a much simpler reformulation suitable for practical applications.

The variables chosen to describe a quantum subsystem and its classical environment are summarized in Table I. In addition, energy and entropy as the generators of reversible and irreversible dynamics, respectively, are listed in this table.

TABLE I.

Variables and the generators of reversible and irreversible dynamics for a quantum subsystem and its classical environment.

VariablesEnergyEntropy
Quantum system ρ on H H(x)ρ kBlnρρ 
Classical environment xM E(xS(x
VariablesEnergyEntropy
Quantum system ρ on H H(x)ρ kBlnρρ 
Classical environment xM E(xS(x

The proper arena for quantum mechanics is provided by separable complete Hilbert spaces, which are complex vector spaces equipped with inner products.31,32 Observables are self-adjoint linear operators on a Hilbert space H. Here, we focus on the evolution of the density matrix ρ, also known as the statistical operator on H. The density matrix characterizes the state of our quantum subsystem, and its time evolution determines the evolution of the averages Aρ=tr(ρA) of all quantum observables A. This perspective corresponds to the Schrödinger picture, which we use throughout this letter.

The discrete, continuous, or mixed set of variables x for the classical environment forms a manifold M. Observables are functions or functionals on the manifold M. For notational simplicity, we assume a discrete set of variables labeled by an index j or k. The modifications required for continuous sets of variables, in particular the proper generalization of matrices and partial derivatives, are explained in detail in Sec. 2 2 and Appendix C of Ref. 33.

Note that we allow the Hamiltonian H(x) to depend on the variables of the environment, thereby introducing a reversible coupling of the quantum system and its environment. The appearance of classical external fields in the Hamiltonian of a quantum system is quite common, for example, a static magnetic field in the Schrödinger equation for discussing Larmor precession or the electromagnetic four-vector potential in the Pauli equation.

At the heart of quantizing the irreversible structure of nonequilibrium thermodynamics is the correlation of two Hilbert space operators A and B29,30,34
(1)
in terms of the dimensionless operator Qα, which is from a set of coupling operators labeled by α. The operators A and B are typically self-adjoint, whereas the coupling operators Qα usually are not (for example, they can be creation and annihilation operators). The second trace term in the definition (1) has been added such that A,Bρuα becomes real, which is crucial for the coupling to a classical system.
The correlation A,Bρuα is closely related to the quantity introduced in Eq. (1) of Ref. 30. Note the symmetry
(2)
and the positivity property
(3)
Moreover, an underlying joint convexity property follows from Lieb’s theorem [see, for example, Eq. (2.120) of Ref. 1].
We further introduce the self-adjoint generalized free-energy operators
(4)
where
(5)
and the matrices Kuα(x) associated with the coupling operators Qα are assumed to be symmetric and positive semidefinite. The simplest dependence of these matrices on u is through a non-negative real prefactor hα(u). We refer to the quantities Fαu(x) as free energy operators because they are combinations of the energy and entropy operators with relative weights proportional to dS and −dEtot.
According to the quantization procedure proposed in Refs. 29 and 30, the evolution of the average Aρ of an observable A of the quantum subsystem is governed by the first-order differential equation
(6)
which actually constitutes the essence of the Dirac-style quantization procedure. The first term expresses reversible evolution, and the second term provides the dissipative coupling to the environment. In the reversible term, we clearly recognize the commutator that, according to Dirac, replaces the Poisson bracket of classical mechanics. The correlation in the second term constitutes the previously suggested replacement for the dissipative bracket in the generalization of Dirac’s quantization procedure.29,30,34
The evolution of the classical environment is given by the first-order differential equations
(7)

The first term describes reversible dynamics of the environment generated by the energy. The energy gradient is multiplied by the antisymmetric Poisson matrix L(x), which is given by the symplectic matrix transformed to non-canonical coordinates. The Poisson bracket of two observables is obtained by multiplying the Poisson matrix from both sides with the gradients of the two arguments of the bracket.

The second term in Eq. (7) describes irreversible dynamics generated by the entropy gradient. The friction matrix M(x) is assumed to be positive-semidefinite so that irreversible dynamics essentially follows the entropy gradient.

The remaining term represents the dissipative coupling between the quantum system and its environment. It is constructed such that the change in the energy of the quantum system, as obtained from Eq. (6) for A = H(x), is compensated by the change in energy of the classical environment, as obtained by multiplying the evolution Eq. (7) with ∂Etot(x)/∂x.

The following degeneracy relations are part of the GENERIC structure of classical nonequilibrium systems:
(8)
They express the conservation of entropy by reversible dynamics and the conservation of energy by irreversible dynamics for any choice of the respective generators of the dynamics.
Equation (6) specifies the evolution of the averages of all quantum observables A evaluated with the density matrix ρ. It can be rewritten as an evolution equation for ρ, which we refer to as the thermodynamic or GENERIC quantum master equation
(9)

Note that the reversible evolution is governed by the commutator with the Hamiltonian, whereas the dissipative contribution possesses a double commutator structure involving coupling and free energy operators. The classical analog of a quantum master equation is a diffusion of the Fokker–Planck equation for the evolution of a probability density, incorporating first and second derivatives for the reversible and irreversible contributions, respectively.

The master Eq. (9) is our fundamental equation for open quantum systems. As a consequence of the definition (1), the second term in Eq. (9) is generally nonlinear in ρ. This nonlinearity of the irreversible contribution is caused by the noncommutativity of quantum observables and implies that, in general, our master equation cannot be of the popular Lindblad form (see, for example, Sec. 3 7 of Ref. 1 for a discussion of nonlinear quantum master equations). The specific conditions under which the GENERIC quantum master equation is linear in ρ are discussed in the  Appendix.

Equations (7) and (9) describe the evolution of the state variables for the classical environment and the quantum system introduced in Table I. They represent the dissipative coupling of a quantum system to a general classical nonequilibrium system as its environment, achieved through Dirac-style quantization of the GENERIC framework. An additional reversible coupling is included in the dependence of the Hamiltonian H(x) of the quantum system on the classical variables x.

With the evolution equation for all system variables at hand, we can now calculate the time-evolution of entropy,
(10)
with the generalized free-energy operators
(11)
Note that these operators fj(x) are closely related to the operators Fαu(x) defined in Eq. (4).

The first term in Eq. (10) describes the entropy production in the classical environment. Note that only the symmetric part of the friction matrix M(x) contributes to entropy production. An antisymmetric contribution to M(x) would describe irreversible processes without entropy production. Historically, this possibility has been introduced by Casimir.19,35 Recent examples of irreversible processes without entropy production include slip phenomena and the energy cascade in turbulence.36 

The second term in Eq. (10) describes the entropy production associated with the dissipative coupling of the quantum system and its classical environment. Note that both contributions to the entropy production (10) are always non-negative. In a thermodynamic setting, this property is more relevant than the complete positivity assumed in Lindblad’s approach.

The first derivation of a nonlinear quantum master equation of the type (9) for a quantum system coupled to a heat bath was achieved by using the projection-operator method.18,25 The same type of nonlinear equation has been recovered from the quantization of GENERIC and illustrated for the examples of a two-level system and a damped harmonic oscillator.34 The zero-temperature limit of the thermodynamic quantum master equation has been discussed in Ref. 37. In addition, heat transport in quantum spin chains has been discussed with this master equation.38 

It has been shown that the nonlinear quantum master Eq. (9) leads to a biexponential decay, a realistic susceptibility profile, and ultralong coherence of a qubit, which is not limited by the energy relaxation time because complete positivity is not imposed.39 It has been recognized that the thermodynamic quantum master equation, which is generally nonlinear, may be of the linear Davies–Lindblad type if the coupling operators Qα are eigenoperators of the Hamiltonian of the quantum subsystem and the associated coupling matrices Kuα(x) are chosen suitably (see  Appendix).30 

The powerful tool of stochastic unravelings1 of quantum master equations for dissipative quantum systems in terms of stochastic jump processes in Hilbert space has been applied to nonlinear equations of the type (9). The nonlinearity can be produced by mean-field interactions in the stochastic jump process.40 One- and two-process unravelings have been developed and tested in Ref. 41.

The dissipative coupling of a quantum system to a time-evolving environment has been explored in Ref. 42. Practical applications include vibrational relaxations in liquids,43 where slower rotational and translational modes can be treated by classical thermodynamics and hydrodynamics, or the Marcus theory of electron transfer in molecular systems,44 where the dielectric environment can be treated by classical thermodynamics and electrodynamics. Further applications include spin-selective radical-ion-pair reactions relevant to photochemistry and photosynthesis,45 quantum dots exchanging energy with two heat baths,46 the coupling of quantum systems to classical opto-electronic systems for the modeling of laser devices,46 semi-classical drift-diffusion-reaction models for the transport of charge carriers in opto-electronic devices,47 and coupled spin dynamics for a sensitivity enhancement of magnetic resonance imaging and spectroscopy.48 

As a final application, we mention that quantum master Eq. (9) provides the foundations of dissipative quantum field theory [cf. Eq. (1.45) of Ref. 49]. In this approach, dissipative smearing regularizes quantum field theory at short distances. Some ontological implications of dissipative quantum field theory have been discussed in Ref. 50. The unraveling of the quantum master equation leads to a new simulation technique for quantum field theory, where the simulation time corresponds to real time. In this context, it has been realized that it is natural to treat interactions as stochastic jumps.

Efficient simulations based on unravelings, in which the interactions of reversible quantum systems are treated as stochastic jumps, have been developed and tested in Refs. 51 and 52. This reformulation naturally leads to a new interpretation of quantum mechanics,53 which is discussed in Sec. III.

It is desirable for an interpretation of quantum mechanics to be based on quantum field theory. For example, in the hydrogen atom, the electron and the proton do not really interact through the static, classical Coulomb potential appearing in the Schrödinger equation, but rather through the exchange of photons.

At any given time, a hydrogen atom consists of an electron, a proton, and a number of photons mediating electromagnetic interactions between the charged particles. As the proton is no longer considered a fundamental particle, one might prefer to say that a hydrogen atom consists of an electron, three quarks, and a number of photons and gluons mediating electromagnetic and strong interactions. In any case, the hydrogen atom has a well-defined content of fundamental particles at any given time. These remarks should clarify that the usual quantum mechanical treatment of the hydrogen atom is a semi-classical approximation involving static classical interactions at a distance. An illustrative toy version of a quantum field theoretical calculation illustrates how bound states can be treated on a more fundamental level.54 The comparison to quantum mechanics is based on the Fourier transformation of the wave functions from position to momentum space.

In contrast to wave functions or density matrices, which describe the properties of an ensemble of hydrogen atoms, stochastic unravelings of density matrices allow us to describe individual atoms. We thus gain a new perspective on both quantum mechanics and quantum technology. The idea that any quantum system has a well-defined particle content suggests that, at any given time, the system can be described by a multiple of a Fock space base vector.49,55,56

Inspired by the discussion of the hydrogen atom, our goal is to reformulate the equations of reversible quantum mechanics in terms of stochastic jump processes, where interactions are treated as discrete collision events. Therefore, we need a splitting of the full Hamiltonian into free and interacting contributions, H = Hfree + Hint. We further assume that there exists a distinguished basis of orthonormal eigenstates |m⟩ of the free Hamiltonian Hfree, which are labeled by the natural number m. The corresponding eigenvalues of Hfree are given by Em. Finally, we assume that the strict superselection rule of quantum field theory is inherited by quantum mechanics, meaning the state of the quantum system at any time t is described by a complex multiple of some base vector |mt⟩. No superpositions between different base vectors are allowed.

With an unraveling in terms of two stochastic processes, ϕt and ψt, in Hilbert space, we wish to reproduce the density matrix ρt evolving according to the von Neumann equation, which is the reversible part of the quantum master Eq. (9) for a quantum system without any dissipative coupling to the environment, by the following expectation evaluated on the probability space of the jump processes:
(12)
where we use Dirac’s bra-ket notation for state vectors (kets) and their duals (bras). The use of the dyadic product in Eq. (12) is motivated by the task of constructing a tensor from the stochastic state vectors in Hilbert space. The expectation E(·) can be thought of as an average over the trajectories of the jump processes.
With the representation (12), the average of a quantum observable A can be obtained as an expectation of stochastic matrix elements,
(13)
This expectation of a bilinear form of stochastic states provides the average of any observable A. Their stochastic nature arises from spontaneous quantum jumps occurring at random times. In the stochastic averaging procedure, nontrivial phase effects and entanglement can arise from this bra-ket formulation. The representation of quantum observables by linear operators on a Hilbert space suggests that there are two sides or aspects associated with every observable A, which, according to Eq. (13), are expressed by the two processes of the unraveling.

The strict superselection rule, which states that linear combinations of different base vectors do not correspond to physical states, reduces the enormous number of possible stochastic jump processes ϕt and ψt. It naturally guides us to a construction of piecewise continuous trajectories with interspersed jumps among basis vectors for the two independent, identically distributed stochastic processes. The following unique stochastic jump process has been constructed in Ref. 53.

1. Free evolution between jumps

If the system between the times t′ and t is represented by a multiple of the base vector m, the complex prefactor oscillates in time and leads to an overall phase shift given by −Em(tt′)/.

2. Random jumps

If the system is represented by a multiple of the base vector m, a positive rate parameter rm characterizes an exponentially decaying probability density for a jump to occur in time. If a jump occurs at time t, a transition from ctm to a new state at time t+ is determined by the following stochastic jump rule:
(14)
To reproduce the von Neumann equation, the rate parameters rm and the weight factors flm have to be chosen such that the following conditions are satisfied:
(15)
The most general solution of these conditions is given by
(16)
and
(17)
where Sm is a positive real parameter, possibly but not necessarily equal to 1. To find a criterion for the choice of the free parameter Sm, we look at the magnitudes of the weight factors flm,
(18)
where Rm is defined by the first equation. If we chose Sm = 1, then Rm > 1 would lead to a total weight factor increasing exponentially in time along any trajectory of the jump process. We therefore prefer to choose Sm < 1 and Rm > 1, fine-tuned such that there occurs no exponential increase or decrease with time and the complex factors flm associated with jumps essentially introduce phase shifts,
(19)
The unique values of Sm and Rm obtained from condition (19) are shown in Fig. 2 as functions of pmm.
FIG. 2.

Magnitude of the weight factors Sm < 1 and Rm > 1 of the stochastic bra-ket unraveling as a function of the probability pmm for self-transitions. As the factor Rm associated with self-transitions is larger than 1, it is convenient to display its inverse. For pmm = 1/2, one finds Sm=1/Rm=520.486<1/2.

FIG. 2.

Magnitude of the weight factors Sm < 1 and Rm > 1 of the stochastic bra-ket unraveling as a function of the probability pmm for self-transitions. As the factor Rm associated with self-transitions is larger than 1, it is convenient to display its inverse. For pmm = 1/2, one finds Sm=1/Rm=520.486<1/2.

Close modal

The stochastic bra-ket formulation of quantum mechanics offers a new interpretation of quantum mechanics. It may be considered an alternative to the currently favored interpretations: Bohmian mechanics,57–61 the GRW approach,62–64 and the many-worlds interpretation.65,66

For reversible quantum systems, the bra and ket processes evolve independently. If the initial conditions are also stochastically independent, the density matrix (12) can be rewritten in the factorized form
(20)
Since the stochastic processes ϕt and ψt are identically distributed, this representation coincides with the density matrix associated with a solution of the Schrödinger equation. Despite the strong superselection rule, the averages E(ϕt) and E(ψt) can be superposition states. These averages do not describe individual pure quantum systems, but they rather represent ensembles of pure quantum systems. The apparent superposition results from stochastic averaging over many individual quantum systems. For mixed states, the factorization (20) does not work.

Note that a constant shift of the Hamiltonian Hint does not affect the von Neumann equation, but as it shifts the matrix elements ⟨m|Hint|m⟩, it affects the jump processes. Therefore, to obtain a unique unraveling, it is important to choose the zero of the interaction energy based on physical arguments.

If there is no natural choice for the zero of energy, one might choose the average energy to be zero. A disadvantage of this choice is that the Hamiltonian depends on the energy of the initial state. An advantage is that steady pure states are described by a time-independent E(ϕt).

In Ref. 53, the usefulness of Eq. (20) is demonstrated in the context of the Einstein–Podolsky–Rosen experiment.67–72 Entanglement arises from the averaging over independent individual states on the bra and ket sides. The wavelike behavior of quantum particles results from an interplay between the bra and ket vectors, as illustrated by the double-slit experiment.53 

We have shown how a theory of quantum dissipation can be developed by quantizing the geometry-based GENERIC framework of nonequilibrium thermodynamics. Thermodynamics is invaluable because it provides a sound language for science and engineering. According to Einstein’s famous appraisal of thermodynamics,73 “It is the only physical theory of universal content, which I am convinced that within the framework of applicability of its basic concepts will never be overthrown (for the special attention of those who are skeptics on principle).”

The geometric structure on which GENERIC is based should, whenever possible, be preserved in developing numerical methods for solving practical problems. For reversible equations, symplectic integrators that preserve the underlying Hamiltonian structure are known to be powerful numerical tools.74,75 For classical dissipative systems, promising initial steps have been taken to reproduce the correct behavior of energy and entropy and to preserve the underlying bracket structure.76–80 For dissipative quantum systems, the development of structure-preserving methods will be even more challenging, particularly when stochastic simulations are included.

The general reversible and irreversible coupling of quantum systems to classical environments is clearly a cornerstone of quantum technology. It is not only a key tool for simplifying or solving problems of practical importance, but it also offers a framework for discussing the measurement problem.

Even closer to the foundations of quantum mechanics is the stochastic bra-ket interpretation described in the second part of this paper. By eliminating the famous paradoxes from quantum mechanics through the application of a strict superselection rule, we may gain deeper intuition about the quantum world. The usual distinction between the classical and quantum worlds is not fundamental but rather a declaration of our lack of understanding and intuition, even after 100 years of quantum mechanics.

According to the bra-ket interpretation, two stochastic jump processes are required to describe an individual quantum system. According to Eq. (13), there are two sides or aspects of quantum variables: a bra and a ket side. The standard equations of quantum mechanics arise after averaging over ensembles of individual systems. The stochastic bra-ket interpretation provides a new quantum reality with a novel implementation of entanglement.

Realism is nice to have for engineers, as well as for down-to-earth scientists and philosophers. In the words often attributed to Max Planck, “When you change the way you look at things, the things you look at change.” An alternative interpretation of quantum mechanics invites new ways of thinking and new questions to be asked, such as: Is a piecewise linear trajectory of a free particle between collisions a valid concept? How close in space and time must the bra and ket trajectories be to contribute to a local measurement? Do particles cease to exist if their bra and ket trajectories move so far apart that they cannot be detected by any local measurement? Should the corresponding loss of particles be compensated by the simultaneous creation of bra and ket vectors in all possible momentum states? Is there a mechanism for bra and ket vectors to stay spatially close, say by favoring momenta in properly selected directions during collisions or by an attractive bra-ket interaction? Can the bra and ket processes describing individual quantum systems be manipulated separately, say by magnetic fields?

A promising tool for addressing these questions is generalized versions of double-slit experiments,81,82 where the electrons of the bra and ket processes always pass through specific slits, but the bra and ket versions of the electrons might pass through different slits.53 By varying slit widths and analyzing high-precision intensity profiles, one could investigate whether interfering spherical waves arise uniformly along slits (according to the Huygens–Fresnel principle) or only by interactions at the edges. With configurations involving more than two slits, possibly arranged in multiple layers, one could try to find out whether the bra and ket electrons pass through well-defined slits or sequences of slits. While time-resolved experiments would provide valuable insights into path lengths, they likely remain beyond current technological capabilities.

The possibility of describing individual quantum systems, rather than ensembles, opens up opportunities for novel applications, although—or perhaps precisely because—these individual systems are subject to the intrinsic randomness of quantum mechanics. This might be relevant in the context of electronic or optical devices that exhibit shot noise. For quantum computers, it might be possible to simulate random variables rather than probability densities. The interplay between the bra and ket aspects of the world may also be key to understanding the transition from particle-wave duality to classical behavior.83 

Advances in the foundations of quantum theory may be valuable for making progress in quantum technology. However, this is by no means a one-way street. The development of quantum devices, conversely, helps us to develop experience with and eventually intuition for quantum mechanics. The challenge of solving urgent practical problems may provide a stronger driving force for progress than the intellectual desire to understand what holds the world together at its core. Curiosity-driven research sometimes leads to curiosities and aberrations. In any case, quantum physics must become classical!

The author has no conflicts to disclose.

Hans Christian Öttinger: Conceptualization (equal); Writing – original draft (equal); Writing – review & editing (equal).

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

Elaborating on the ideas of Ref. 30, we here determine the conditions under which the GENERIC master Eq. (9) becomes linear. The identity
(A1)
when used in the definition (1), leads to
(A2)
For some weight function h(u), after an integration by parts, we obtain
(A3)
The nonlinear integral part of the entropy-generated contribution must be canceled by the energy-generated contribution
(A4)
To see the conditions for cancellation more clearly, we rewrite the generalized free-energy operators (4) as
(A5)
with the “temperature”
(A6)
If we want Tα(x) to be independent of u for all choices of the generators Etot(x) and S(x) of reversible and irreversible dynamics of the classical environment, Kαu(x) must depend on an overall factor of u that cancels out in the definition (A6). For a cancellation of the integral terms in Eqs. (A3) and (A4) to arise, the dependence of Kαu(x) on u must be exponential,
(A7)
If we further assume that the operators Qα are eigenoperators of the Hamiltonian H(x)
(A8)
the condition for the cancellation of the nonlinear integral terms in the GENERIC quantum master equation becomes
(A9)
If the Hamiltonian H(x) actually depends on x, we also expect the coupling operators and their frequencies introduced in Eq. (A8) to depend on x. The final linear quantum master equation of the GENERIC type is given by the linear part of Eq. (A3),
(A10)
Note that for
(A11)
the α contribution to dissipation in Eq. (A10) vanishes, as can be verified by using
(A12)
1.
H.-P.
Breuer
and
F.
Petruccione
,
The Theory of Open Quantum Systems
(
Oxford University Press
,
Oxford
,
2002
).
2.
U.
Weiss
,
Quantum Dissipative Systems
,
Series in Modern Condensed Matter Physics
, 3rd ed. (
World Scientific
,
Singapore
,
2008
), Vol.
13
.
3.
P. K.
Feyerabend
, “
Problems of microphysics
,” in
Frontiers of Science and Philosophy, University of Pittsburgh Series in the Philosophy of Science No. 1
, edited by
R. G.
Colodny
(
University of Pittsburgh Press
,
Pittsburgh
,
1962
), pp.
189
283
.
4.
D. Z.
Albert
,
Quantum Mechanics and Experience
(
Harvard University Press
,
Cambridge, MA
,
1992
).
5.
T.
Maudlin
, “
Three measurement problems
,”
Topoi
14
,
7
15
(
1995
).
6.
J. A.
Barrett
, “
Entanglement and disentanglement in relativistic quantum mechanics
,”
Stud. Hist. Philos. Sci. B
48
,
168
174
(
2014
).
7.
R. P.
Feynman
, “
Simulating physics with computers
,”
Int. J. Theor. Phys.
21
,
467
488
(
1982
).
8.
L.
Boltzmann
, “
Weitere Studien über das Wärmegleichgewicht unter Gasmolekülen
,”
Wien. Ber.
66
,
275
370
(
1872
).
9.
H.
Nyquist
, “
Thermal agitation of electric charge in conductors
,”
Phys. Rev.
32
,
110
113
(
1928
).
10.
H. B.
Callen
and
T. A.
Welton
, “
Irreversibility and generalized noise
,”
Phys. Rev.
83
,
34
40
(
1951
).
11.
R.
Kubo
,
M.
Toda
, and
N.
Hashitsume
,
Nonequilibrium Statistical Mechanics
,
Statistical Physics
, 2nd ed. (
Springer
,
Berlin
,
1991
), Vol.
2
.
12.
H. C.
Öttinger
,
M.
Peletier
, and
A.
Montefusco
, “
A framework of nonequilibrium statistical mechanics. I. Role and types of fluctuations
,”
J. Non-Equilib. Thermodyn.
46
,
1
13
(
2021
).
13.
A.
Montefusco
,
M.
Peletier
, and
H. C.
Öttinger
, “
A framework of nonequilibrium statistical mechanics. II. Coarse-graining
,”
J. Non-Equilib. Thermodyn.
46
,
15
33
(
2021
).
14.
R.
Zwanzig
, “
Memory effects in irreversible thermodynamics
,”
Phys. Rev.
124
,
983
992
(
1961
).
15.
H.
Mori
, “
Transport, collective motion, and Brownian motion
,”
Prog. Theor. Phys.
33
,
423
455
(
1965
).
16.
H.
Mori
, “
A continued-fraction representation of the time-correlation functions
,”
Prog. Theor. Phys.
34
,
399
416
(
1965
).
17.
B.
Robertson
, “
Equations of motion in nonequilibrium statistical mechanics
,”
Phys. Rev.
144
,
151
161
(
1966
).
18.
H.
Grabert
,
Projection Operator Techniques in Nonequilibrium Statistical Mechanics
(
Springer
,
Berlin
,
1982
).
19.
S. R.
de Groot
and
P.
Mazur
,
Non-Equilibrium Thermodynamics
, 2nd ed. (
Dover
,
New York
,
1984
).
20.
M.
Grmela
, “
Particle and bracket formulations of kinetic equations
,” in
Fluids and Plasmas: Geometry and Dynamics, Contemporary Mathematics No. 28
, edited by
J. E.
Marsden
(
American Mathematical Society
,
Providence, Rhodes Island
,
1984
), pp.
125
132
.
21.
M.
Grmela
, “
Bracket formulation of dissipative fluid mechanics equations
,”
Phys. Lett. A
102
,
355
358
(
1984
).
22.
M.
Grmela
and
H. C.
Öttinger
, “
Dynamics and thermodynamics of complex fluids. I. Development of a general formalism
,”
Phys. Rev. E
56
,
6620
6632
(
1997
).
23.
H. C.
Öttinger
and
M.
Grmela
, “
Dynamics and thermodynamics of complex fluids. II. Illustrations of a general formalism
,”
Phys. Rev. E
56
,
6633
6655
(
1997
).
24.
G.
Lindblad
, “
On the generators of quantum dynamical semigroups
,”
Commun. Math. Phys.
48
,
119
130
(
1976
).
25.
H.
Grabert
, “
Nonlinear relaxation and fluctuations of damped quantum systems
,”
Z. Phys. B: Condens. Matter
49
,
161
172
(
1982
).
26.
M.
Esposito
,
K.
Lindenberg
, and
C.
Van den Broeck
, “
Entropy production as correlation between system and reservoir
,”
New J. Phys.
12
,
013013
(
2010
).
27.
Thermodynamics in the Quantum Regime: Fundamental Aspects and New Directions, Fundamental Theories of Physics No. 195
, edited by
F.
Binder
,
L. A.
Correa
,
C.
Gogolin
,
J.
Anders
and
G.
Adesso
(
Springer
,
Cham
,
2018
).
28.
G. T.
Landi
and
M.
Paternostro
, “
Irreversible entropy production: From classical to quantum
,”
Rev. Mod. Phys.
93
,
035008
(
2021
).
29.
H. C.
Öttinger
, “
The geometry and thermodynamics of dissipative quantum systems
,”
Europhys. Lett.
94
,
10006
(
2011
).
30.
D.
Taj
and
H. C.
Öttinger
, “
Natural approach to quantum dissipation
,”
Phys. Rev. A
92
,
062128
(
2015
).
31.
P. A. M.
Dirac
,
The Principles of Quantum Mechanics
,
International Series of Monographs on Physics
, 4th ed. (
Oxford University Press
,
Oxford
,
1981
), Vol.
27
.
32.
L. I.
Schiff
,
Quantum mechanics
,
International Series in Pure and Applied Physics
, 3rd ed. (
McGraw-Hill
,
New York
,
1968
).
33.
H. C.
Öttinger
,
Beyond Equilibrium Thermodynamics
(
Wiley
,
Hoboken
,
2005
).
34.
H. C.
Öttinger
, “
Nonlinear thermodynamic quantum master equation: Properties and examples
,”
Phys. Rev. A
82
,
052119
(
2010
).
35.
H. B. G.
Casimir
, “
On Onsager’s principle of microscopic reversibility
,”
Rev. Mod. Phys.
17
,
343
350
(
1945
).
36.
H. C.
Öttinger
, “
Irreversible dynamics, Onsager–Casimir symmetry, and an application to turbulence
,”
Phys. Rev. E
90
,
042121
(
2014
).
37.
H. C.
Öttinger
, “
Zero-temperature limit of thermodynamic quantum master equations
,”
Phys. Rev. A
98
,
012131
(
2018
).
38.
M.
Borrelli
and
H. C.
Öttinger
, “
Dissipation in spin chains using quantized nonequilibrium thermodynamics
,”
Phys. Rev. A
106
,
022220
(
2022
).
39.
J.
Flakowski
,
M.
Osmanov
,
D.
Taj
, and
H. C.
Öttinger
, “
Biexponential decay and ultralong coherence of a qubit
,”
Europhys. Lett.
113
,
40003
(
2016
).
40.
H. C.
Öttinger
, “
Stochastic process behind nonlinear thermodynamic quantum master equation. I. Mean-field construction
,”
Phys. Rev. A
86
,
032101
(
2012
).
41.
J.
Flakowski
,
M.
Schweizer
, and
H. C.
Öttinger
, “
Stochastic process behind nonlinear thermodynamic quantum master equation. II. Simulation
,”
Phys. Rev. A
86
,
032102
(
2012
).
42.
M.
Osmanov
and
H. C.
Öttinger
, “
Open quantum systems coupled to time-dependent classical environments
,”
Int. J. Thermodyn.
34
,
1255
1264
(
2013
).
43.
D. W.
Oxtoby
, “
Vibrational relaxation in liquids
,”
Annu. Rev. Phys. Chem.
32
,
77
101
(
1981
).
44.
V.
May
and
O.
Kühn
,
Charge and Energy Transfer Dynamics in Molecular Systems
(
Wiley-VCH
,
Weinheim
,
2011
).
45.
I. K.
Kominis
, “
Radical-ion-pair reactions are the biochemical equivalent of the optical double-slit experiment
,”
Phys. Rev. E
83
,
056118
(
2011
).
46.
A.
Mielke
, “
On thermodynamical couplings of quantum mechanics and macroscopic systems
,” in
Mathematical Results in Quantum Mechanics
, edited by
P.
Exner
,
W.
König
and
H.
Neidhardt
(
World Scientific
,
Singapore
,
2015
), pp.
331
348
.
47.
M.
Kantner
,
A.
Mielke
,
M.
Mittnenzweig
, and
N.
Rotundo
, “
Mathematical modeling of semiconductors: From quantum mechanics to devices
,” in
Topics in Applied Analysis and Optimisation, CIM Series in Mathematical Sciences
, edited by
M.
Hintermüller
and
J. F.
Rodrigues
(
Springer
,
Cham
,
2019
), pp.
269
293
.
48.
R.
Levi
,
S.
Masis
, and
E.
Buks
, “
Instability in the Hartmann–Hahn double resonance
,”
Phys. Rev. A
102
,
053516
(
2020
).
49.
H. C.
Öttinger
,
A Philosophical Approach to Quantum Field Theory
(
Cambridge University Press
,
Cambridge
,
2017
).
50.
A.
Oldofredi
and
H. C.
Öttinger
, “
The dissipative approach to quantum field theory: Conceptual foundations and ontological implications
,”
Eur. J. Philos. Sci.
11
,
18
(
2021
).
51.
R.
Chessex
,
M.
Borrelli
, and
H. C.
Öttinger
, “
Fixed-point quantum Monte Carlo method: A combination of density-matrix quantum Monte Carlo method and stochastic unravellings
,”
Phys. Rev. A
105
,
062803
(
2022
).
52.
R.
Chessex
,
M.
Borrelli
, and
H. C.
Öttinger
, “
Dynamical triplet unraveling: A quantum Monte Carlo algorithm for reversible dynamics
,”
Phys. Rev. A
106
,
022222
(
2022
).
53.
H. C.
Öttinger
, “
Stochastic bra-ket interpretation of quantum mechanics
,”
J. Phys. Commun.
8
,
105004
(
2024
).
54.
T.
Pashby
and
H. C.
Öttinger
,
“Quantum jump revival” (in press)
(
2025
).
55.
V.
Fock
, “
Konfigurationsraum und zweite Quantelung
,”
Z. Phys.
75
,
622
647
(
1932
).
56.
P.
Teller
,
An Interpretive Introduction to Quantum Field Theory
(
Princeton University Press
,
Princeton
,
1995
).
57.
D.
Bohm
, “
A suggested interpretation of the quantum theory in terms of ‘hidden’ variables
,”
Phys. Rev.
55
,
166
179
(
1952
).
58.
D.
Dürr
,
S.
Goldstein
, and
N.
Zanghí
, “
Quantum equilibrium and the origin of absolute uncertainty
,”
J. Stat. Phys.
67
,
843
907
(
1992
).
59.
M.
Esfeld
,
M.
Hubert
,
D.
Lazarovici
, and
D.
Dürr
, “
The ontology of Bohmian mechanics
,”
Br. J. Philos. Sci.
65
,
773
796
(
2014
).
60.
D.-A.
Deckert
,
M.
Esfeld
, and
A.
Oldofredi
, “
A persistent particle ontology for quantum field theory in terms of the Dirac sea
,”
Br. J. Philos. Sci.
70
,
747
770
(
2019
).
61.
O.
Passon
,
Bohmsche Mechanik
, 2nd ed. (
Harri Deutsch
,
Frankfurt
,
2010
).
62.
G. C.
Ghirardi
,
A.
Rimini
, and
T.
Weber
, “
Unified dynamics for microscopic and macroscopic systems
,”
Phys. Rev. D
34
,
470
491
(
1986
).
63.
J. S.
Bell
, “
Are there quantum jumps?
,” in
Schrödinger, Centenary Celebration of a Polymath
, edited by
C. W.
Kilmister
(
Cambridge University Press
,
Cambridge
,
1987
), pp.
41
52
.
64.
R.
Tumulka
, “
A relativistic version of the Ghirardi–Rimini–Weber model
,”
J. Stat. Phys.
125
,
825
844
(
2006
).
65.
H.
Everett
III
, “‘
Relative state’ formulation of quantum mechanics
,”
Rev. Mod. Phys.
29
,
454
462
(
1957
).
66.
D.
Wallace
,
The Emergent Multiverse: Quantum Theory According to the Everett Interpretation
(
Oxford University Press
,
Oxford
,
2012
).
67.
A.
Einstein
,
B.
Podolsky
, and
N.
Rosen
, “
Can quantum-mechanical description of physical reality be considered complete?
,”
Phys. Rev.
47
,
777
780
(
1935
).
68.
J. S.
Bell
, “
On the Einstein Podolsky Rosen paradox
,”
Phys. Phys. Fiz.
1
,
195
200
(
1964
).
69.
J. S.
Bell
, “
On the problem of hidden variables in quantum mechanics
,”
Rev. Mod. Phys.
38
,
447
452
(
1966
).
70.
S. J.
Freedman
and
J. F.
Clauser
, “
Experimental test of local hidden-variable theories
,”
Phys. Rev. Lett.
28
,
938
941
(
1972
).
71.
A.
Aspect
,
P.
Grangier
, and
G.
Roger
, “
Experimental realization of Einstein–Podolsky–Rosen–Bohm gedankenexperiment: A new violation of Bell’s inequalities
,”
Phys. Rev. Lett.
49
,
91
94
(
1982
).
72.
A.
Aspect
, “
Bell’s theorem: The naive view of an experimentalist
,” in
Quantum (Un)speakables: From Bell to Quantum Information
, edited by
R. A.
Bertlmann
and
A.
Zeilinger
(
Springer
,
Berlin
,
2002
), pp.
119
153
.
73.
A.
Einstein
,
Autobiographical Notes
, in
Albert Einstein: Philosopher–Scientist, the Library of Living Philosophers No. VII
, edited by
P. A.
Schilpp
(
The Library of Living Philosophers, Inc.
,
Evanston
,
1949
), pp.
1
96
.
74.
J.
Moser
,
Lectures on Hamiltonian Systems
(
Memoirs of the American Mathematical Society
,
1968
), Vol.
81
, pp.
1
60
.
75.
J. M.
Sanz-Serna
and
M. P.
Calvo
,
Numerical Hamiltonian Problems, Applied Mathematics and Mathematical Computation
(
Chapman & Hall
,
London
,
1994
), Vol.
7
.
76.
M.
Krüger
,
M.
Groß
, and
P.
Betsch
, “
An energy-entropy-consistent time stepping scheme for nonlinear thermo-viscoelastic continua
,”
Z. Angew. Math. Mech.
96
,
141
178
(
2016
).
77.
D.
Portillo
,
J. C.
García Orden
, and
I.
Romero
, “
Energy-entropy-momentum integration schemes for general discrete non-smooth dissipative problems in thermomechanics
,”
Int. J. Numer. Methods Eng.
112
,
776
802
(
2017
).
78.
M.
Kraus
and
E.
Hirvijoki
, “
Metriplectic integrators for the Landau collision operator
,”
Phys. Plasmas
24
,
102311
(
2017
).
79.
H. C.
Öttinger
, “
GENERIC integrators: Structure preserving time integration for thermodynamic systems
,”
J. Non-Equilib. Thermodyn.
43
,
89
100
(
2018
).
80.
X.
Shang
and
H. C.
Öttinger
, “
Structure-preserving integrators for dissipative systems based on reversible-irreversible splitting
,”
Proc. R. Soc. A
476
,
20190446
(
2020
).
81.
A.
Zeilinger
,
R.
Gähler
,
C. G.
Shull
,
W.
Treimer
, and
W.
Mampe
, “
Single- and double-slit diffraction of neutrons
,”
Rev. Mod. Phys.
60
,
1067
1073
(
1988
).
82.
Y.
Aharonov
,
E.
Cohen
,
F.
Colombo
,
T.
Landsberger
,
I.
Sabadini
,
D. C.
Struppa
, and
J.
Tollaksen
, “
Finally making sense of the double-slit experiment
,”
Proc. Natl. Acad. Sci. U. S. A.
114
,
6480
6485
(
2017
).
83.
S. V.
Gantsevich
, “
Common sense and quantum mechanics
,”
Ann. Math. Phys.
5
,
196
198
(
2022
).