1. Introduction
Consider a global Hilbert space , which contains all single-particle degrees of freedom available to the universe. From this we construct the Fock space , within which all many-body theories are described. In the study of open quantum systems, one typically divides the universe into complementary sets of orthogonal, single-particle states describing the subsystem of interest and a thermal reservoir . The Fock space then decomposes as the familiar product .
Denote by
the set of linear operators acting on
. Any observable
may, of course, be written in the form
, where
and
describes the coupling between the two subsystems. One may then ask whether the expectation value
can be sensibly divided into system and reservoir contributions. In particular, what fraction of the averaged coupling term
should be assigned to each subsystem? For a partitioning of
into subsystems of comparable size it seems intuitive that the coupling should be partitioned symmetrically, assigning
to each subsystem. However, in the context of the thermodynamics of open quantum systems, an answer to this question has remained elusive, due to subtleties in the distinction between system internal energy and heat [
1,
2,
3,
4,
5,
6,
7,
8]. This thermodynamic bookkeeping typically falls into one of two camps: either half of the coupling Hamiltonian is assigned to the system, as in the symmetric partition, or all of it is. The remaining fraction is then assigned to the reservoir and used to describe heat dissipated.
In this article, we maintain that any partition of the energetics between two subsystems should be inherited from the division of the single-particle states . Moreover, it will be shown that such a “Hilbert-space partition" of any observable leads directly to the symmetric partition.
Of central interest in the study of open quantum systems is the von Neumann entropy
, where
is the density matrix
1 describing the state of the global system in
. A description of the
system entropy may be constructed by means of the reduced density matrix, defined from
by tracing out the reservoir degrees of freedom,
. The reduced entropy of the system is then defined as
. Using a quantum master equation to describe the dynamics of the reduced density matrix for the system, it has been shown [
9,
10] that the reduced entropy obeys
where
is a heat current flowing between the system and reservoir and
is the entanglement entropy between the subsystems. While interesting from the point of view of quantum information theory, this production of entanglement entropy, we argue, is not thermodynamic in character.
As a simple illustration of our argument, consider an infinite, fermionic, tight-binding chain in thermal equilibrium at temperature
T.
Here we have taken the chemical potential to lie in the center of the electron band. With the chain in thermal equilibrium, one might expect that all extensive quantities should be uniformly distributed throughout the chain. In particular, for the entropy of a single site we anticipate that
, where
N is the number of sites in the chain and
with
the Fermi-Dirac distribution for orbital
at temperature
. On the other hand, the reduced state for a single site will be of the form
where
is the number operator for the site and
f gives the probability that this site is occupied. Using the results presented in Ref. [
11], it can be shown that
which is to be expected for the occupancy of a single site given
in the center of the band, regardless of temperature. The reduced entropy is thus
independent of temperature. Such an entropy is to be expected for the chain’s classical counterpart in a microcanonical ensemble, with
N sites filled by
particles. The classical picture neglects the entanglement between the fermions occupying the chain; though non-interacting, the antisymmetric nature of the fermionic wavefunctions enforces a maximal entanglement for all states in the ensemble. Indeed, upon comparing stot and Sred, we see that
, with
the total entanglement entropy of the lattice sites.
In particular, as
the chain approaches a pure state with all particles maximally entangled. Thus, we should have that
, in accordance with the Third Law of Thermodynamics. A comparison between the entropy density
and reduced entropy is shown in
Figure 1.
Clearly, the reduced entropy does not have a meaningful thermodynamic interpretation, whereas the entropy density does, at least for this toy model. In what follows, it will be shown that the Hilbert-space partition can be employed to give a local description of the entropy which does have a meaningful thermodynamic interpretation. Importantly, it will be shown that this partitioned entropy does not measure the entanglement between the subsystems.
2. Hilbert-Space Partition
In what follows, we will denote by the partition of observable over the subspace . We seek a partition which is
Inherited from the division of single-particle states on , and
Additive over subspaces: .
For the sake of clarity, we focus for now on one-body observables
, and discuss the generalizations to
N-body observables in
Section 6. The matrix elements
can be extracted in the form of an observable in first quantization,
, which acts on the single-particle Hilbert space. As it is this space which is being divided into system and reservoir states, the partition should be constructed at the level of these matrix elements. By second quantization of the partitioned operator
we recover
. This construction is outlined in
Figure 2.
For the partition of
O, we define
where
is an orthogonal projector onto
and the anti-commutator
is included to ensure that
is Hermitian. The Fock-space operator
is then given by
That this partition satisfies the additivity condition is a simple consequence of the fact that , so that .
Furthermore, defining
we see that
. Performing second quantization, one therefore finds that this partition is equivalent to the symmetric partition, wherein the system-reservoir coupling is partitioned equally between subsystems
In the remainder of this paper, we will use to denote the second quantization of an operator .
3. Time Dependence
In order to construct the local dynamics of a partitioned expectation value , it will be useful to consider the Heisenberg evolution of the partitioned operators themselves. In what follows, we consider only non-interacting theories. Even in this simple case there is, however, some ambiguity in constructing this evolution: Should one first evolve the operator forward in time, and then construct the partition, ? Or partition and evolve this forward in time, ? Contrary to what one may expect, the two approaches are not generally equivalent.
Proposition 1. if and only if .
Proof. In the absence of interactions, we have in general that
. Where
is the usual evolution operator generated by Hamiltonian
. Then the partition of the time-evolved operator is
with
the Heisenberg evolution of
O. On the other hand, the time evolution of the partitioned operator gives
Clearly, then, these two expressions will only be equivalent if
so that
. □
In order to preserve the equivalence between the Schrödinger and Heisenberg pictures, one must use Ooft to describe the dynamics of the partitioned observable. We emphasize that in general , and so contrary to what one may expect, the evolution of the partition is not given by the partitioning of .
The operator
therefore obeys the inhomogeneous continuity equation
With
In the above, the time dependence of the projection operator
has been suppressed. Using the fact that the second quantization of a commutator is the commutator of the second quantized operators, it can readily be shown that
, where
is the number operator for subsystem
, implying that
is a probability current operator. Thus, we interpret
as a current operator for transport of
into the subspace
. Sigma can be written
, and so we interpret this term as the local production of
within the subspace
. The Fock-space operator
therefore obeys a similar continuity equation
As an illustrative application of this partition, consider the energy current passing through site
in the tight-binding model
where the sum on
denotes a sum over nearest neighbors. Using JA, we find for the energy current
which implies the textbook definition [
12] for the energy current operator:
4. Density of One-Body Observables
Up until now, we have discussed only partitions over a discrete subspace of
. The same construction can be applied also to subsets
by replacing the projector
with the operator
for
. We thus define the density of a one-body observable
in much the same way as before:
and
Note that, as with the discrete partition, need not be diagonal in position representation to define . Rather, we argue that describes the influence of the non-local observable at the location .
This interpretation may be clarified by considering the operator
in position representation
where
are the usual fermionic field creation and annihilation operators, and
. The density in hdensity can be rewritten in terms of
as
Following the same arguments as in the previous section, we find that the density operator obeys a continuity equation
where
In the above,
is the usual momentum operator. Moreover, in slight contrast to J,Sigma, we have made explicit the time evolution of the projection operator
. The definitions of current density in jO,j are consequences of the following claim:
Proposition 2. , where .
Proof. Only the kinetic part of the Hamiltonian is non-vanishing in the commutator, since for any two states
and
assuming that
V acts as a simple multiplication operator in position representation. Then
To proceed, we consider the matrix elements
Therefore, for any two states
and
,
, which proves the claim. □
Note that gives the usual probability current density of the state .
5. Entropy Partitions
Of particular interest is the partition of the von Neumann entropy. Define in the Schrödinger picture the entropy operator
where
is the density matrix for the global ensemble. Then
due to the unitary evolution of
. The Heisenberg picture entropy operator is then
. So the entropy operator is constant, in agreement with our expectation that the global entropy should be constant under unitary evolution.
In the absence of inter-particle interactions, the state of the system may be taken to be a product of the form
where
is any set of single-particle orbitals which span
and
describes the probability that the state
is occupied. We define the
statistical basis to be the set of single-particle orbitals over which the density matrix factorizes. Such a state may describe, for example, a quantum system initially in equilibrium that is subsequently acted upon by a time-dependent external force.
For such a product state, the entropy operator becomes a sum of particle- and hole-ordered single-particle operators
The arguments of the previous sections hold just as well for a hole-ordered operator, and so the entropy may be partitioned as in partition,hatpartition. In particular, its partition and density obey the continuity equations
where
are the entropy density operator and the entropy current density operator, respectively. The net entropy current operator
can be obtained as minus the surface integral of Eq. (
36).
Crucially, as a consequence of global entropy conservation under unitary evolution, we find that there is no local entropy production in Scont,Scontdens. Comparing to Sredcons, we interpret this to mean that the entropy partition proposed in this article does not measure entanglement entropy between subsystems. This entropy partition may therefore provide a more faithful description of the local thermodynamic entropy of the system [
13].
5.1. Evolution of Entropy Density
In this section, we illustrate the continuity equation [Scont] by modeling the evolution of the entropy density in a one-dimensional chain of fermions in the tight-binding model. We consider a finite chain of
sites with zero occupancy and a single, isolated site with a probability of occupancy of
which are coupled on the left end of the chain at time
. The initial density matrix for this configuration is
where
(
) creates (annihilates) a fermion on the initially isolated site and
are eigenstates of the uncoupled chain. The unoccupied sites contribute nothing to the total entropy, and so the entropy operator is
Using Ooft with projection operator
the entropy density operator for lattice site
m may be expressed in the statistical basis (see discussion following rho) as
The entropy density on site
m is therefore given as
with evolution governed by the tight-binding Hamiltonian on the coupled lattice
Here is the nearest-neighbor hopping integral.
From JA,J we obtain the following expression for the entropy current density into site
m
In
Figure 3, we plot the entropy density and entropy current density on the initially uncoupled site (the 0
th site) as a function of time. The entropy density,
, is calculated using Sm, and the entropy current density using JSm. In the same figure, we also plot the difference
in order to emphasize the local conservation of entropy. The timescale
is defined as the time required for a wavepacket traveling at the maximum group velocity to traverse the full lattice.
The evolution of the entropy density on the lattice is explored further in
Figure 4,
Figure 5.
Figure 4 shows the entropy density evolution on the first ten lattice sites shortly after the 0
th site is coupled to the chain.
Figure 5 shows the entropy density on the full lattice over the full timescale
. Note the change in vertical scale in
Figure 5 from
to
. Initially, only the uncoupled site has entropy
. As the system evolves, this entropy is spread out over the rest of the lattice while maintaining a constant total entropy of
.
5.2. Consistency with the Third Law
Applying the Hilbert space partition to the entropy operator leads to a system entropy operator
In light of the above discussion, this would imply that the heat added to an open quantum system should be defined as [
13]
for linear deviations from equilibrium, where
is the system number operator and
is the external work done on the system.
Various authors [
1,
2,
3,
4,
6,
7,
8] have suggested including different fractions of the coupling energy
in the heat partition. However, it can be shown that any partition of the entropy apart from Spart, corresponding to the symmetric partition advocated in Refs. [
1,
3,
4,
13], leads to a violation of the Third Law of Thermodynamics. Define the
-partition of
as
Proposition 3. Let be a product state of the form rho. Then for any , diverges in the limit for any k.
Proof. Define the operators in first quantization and
so that with and being their corresponding particle- and hole-like quantizations. We will focus our attention on , anticipating that the treatment for will be identical, and drop the superscripts.
The coupling term can therefore be written as the second quantization of the operator
If we write
and make use of
, then the last line becomes
Consider now the matrix elements in the statistical basis,
2:
where
. We find, then, for the partitioned entropy
The first term in Eq. (
50) diverges if any
, unless
or
. The latter condition applies only to a partition of the statistical basis, and is certainly not true in general. Thus, we conclude that for a general partition of the entropy of a partially pure state,
3 diverges if
. □
From eq5 we conclude that the only well defined partition is , the Hilbert-space partition. In particular, this implies that the partitioned entropy of the Gibbs state diverges in the limit if , since all states above the Fermi level will be unoccupied with definite probability. We conclude that the Hilbert-space partition is the only thermodynamically consistent partition of the entropy and heat for an open quantum system in equilibrium with its surroundings.
Note, though, that our conclusion is even stronger than this. From wide it is clear that the partitioned entropy will diverge for if there are any localized pure states in the product ensemble rho.
6. Many-Body Operators
Until now, we have considered only partitions for one-body observables. It is, however, also possible to construct a partition for N-body observables based on the partition of the single-particle Hilbert space. In the remainder of this section, we generalize the above partitions first to body, and then N-body observables.
6.1. Two-Body Operators
Consider first a two-body operator
which acts on the Hilbert space
. We wish to generalize the partition of
O onto a subspace
in such a way that
, where
This can be accomplished by adjusting the projection operator as follows
:
Then, because
and
,
as needed. It then follows that the partition of the second quantized operator
is
One may generalize the density of observables in the same manner, from which we obtain, in position representation,
As motivation for this construction, consider the Coulomb interaction
where
and
denotes the normal ordering of
. Defining
we see that
. rho2 then implies that the energy density is
in analogy with the classical description.
6.2. N-Body Observables
For the matrix elements of an
N-body operator
we define the partition
Proposition 4.
For any N-body observable , its partition over is given by
The proof follows straightforwardly from an application of Wick’s theorem to the anticommutator . Upon comparing the normal ordering to the second quantization of ON, one can see that they differ by exactly . A detailed proof is given in appendix A.
Corollary 1. The expectation value of a partitioned many-body observable obeys the continuity equation in contflux.
A detailed proof is given in appendix B.
Following the same arguments as in appendix A, one can show that the density of many-body observables is similarly defined.
Proposition 5.
The density of any N-body observable may be defined as
7. Conclusions
In this article, we have developed a framework to partition quantum observables based on a partition of the underlying single-particle Hilbert space. We have provided explicit expressions for the partition of Fock-space operators corresponding to generic
N-body observables. For the case of an open quantum system, a bipartite partition between system (S) and reservoir (R) was applied to both the Hamiltonian and the entropy of the system. The Hilbert-space partition was shown to correspond to a symmetric (
) partition [
1,
3,
4,
13] of the off-diagonal 1-body observables, such as the coupling Hamiltonian
between system and reservoir. Any partition with
was shown to lead to a singular entropy, and hence does not provide a basis to construct a consistent thermodynamics from the statistical mechanics of the problem.
Funding
This work was funded in part by the U.S. Department of Energy (DOE), Office of Science, under Award No. DE-SC0006699.
Acknowledgments
We thank Carter Eckel, Ferdinand Evers, Marco Jimenez, Parth Kumar, and Yiheng Xu for useful discussions.
Conflicts of Interest
The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.
Appendix A. Proof of best
The proof will require two lemmas. Define by
the map which second quantizes an operator
.
Lemma A1.
Let A and B be two one-body operators. Then
Proof.
Similarly,
. The statement of the proof follows immediately from here. □
Using the fact that it is readily apparent from this first lemma that the partition of any one-body observable can be constructed as in best. To show that this holds for any body observable we will need the following.
Lemma A2.
Let
be a normal ordered body operator and be a one-body operator. Then
Proof. From Wick’s Theorem
and similarly
The rest follows straightforwardly from here. □
While it seems reasonable that 1b may be extended to
body observables, one needs to be careful with the normalization. From L2 with
,
and so
Therefore
confirming that the additional factor of
in best correctly normalizes the partition.
Finally, it remains to be shown that best follows from the partition defined in ON.
Proof. Using L2 with
we recognize the first term in brackets in l2 as the matrix element
with
the identity on
, the
particle subspace. Because
we conclude that
Therefore
□
Appendix B. N-Body Continuity Equation
Here, we show that a continuity equation for the expectation value of the partition holds for an
body observable
. In the Schrödinger picture
Above, we allow for explicit time dependence in
and
, and the evolution of the system is accounted for in the density matrix. To prove this we shift contflux into the Schrödinger picture and then apply equation best. First, recall that for any operator
in the Schrödinger picture
Using the identity
we have that
Then, because the commutator of any
body and one-body operator is a normal ordered
body operator,
]. Therefore, in the absence of interactions,
claim then follows immediately from the definition of a partition of an
N-body observable, best. Note that this also provides an extension of the continuity equation for time-dependent Hamiltonians.
References
- Ludovico, M.F.; Lim, J.S.; Moskalets, M.; Arrachea, L.; Sánchez, D. Dynamical energy transfer in ac-driven quantum systems. Phys. Rev. B 2014, 89, 161306. [Google Scholar] [CrossRef]
- Esposito, M.; Ochoa, M.A.; Galperin, M. Nature of heat in strongly coupled open quantum systems. Phys. Rev. B 2015, 92, 235440. [Google Scholar] [CrossRef]
- Bruch, A.; Thomas, M.; Viola Kusminskiy, S.; von Oppen, F.; Nitzan, A. Quantum Thermodynamics of the Driven Resonant Level Model. Physical Review B 2016, 93, 115318. [Google Scholar] [CrossRef]
- Haughian, P.; Esposito, M.; Schmidt, T.L. Quantum Thermodynamics of the Resonant-Level Model with Driven System-Bath Coupling. Physical Review B 2018, 97, 085435. [Google Scholar] [CrossRef]
- Talkner, P.; Hänggi, P. Colloquium : Statistical Mechanics and Thermodynamics at Strong Coupling: Quantum and Classical. Reviews of Modern Physics 2020, 92, 041002. [Google Scholar] [CrossRef]
- Strasberg, P.; Winter, A. First and Second Law of Quantum Thermodynamics: A Consistent Derivation Based on a Microscopic Definition of Entropy. PRX Quantum 2021, 2, 030202. [Google Scholar] [CrossRef]
- Bergmann, N.; Galperin, M. A Green’s Function Perspective on the Nonequilibrium Thermodynamics of Open Quantum Systems Strongly Coupled to Baths: Nonequilibrium Quantum Thermodynamics. The European Physical Journal Special Topics 2021, 230, 859–866. [Google Scholar] [CrossRef]
- Lacerda, A.M.; Purkayastha, A.; Kewming, M.; Landi, G.T.; Goold, J. Quantum Thermodynamics with Fast Driving and Strong Coupling via the Mesoscopic Leads Approach. Physical Review B 2023, 107, 195117. [Google Scholar] [CrossRef]
- Esposito, M.; Lindenberg, K.; den Broeck, C.V. Entropy production as correlation between system and reservoir. New Journal of Physics 2010, 12, 013013. [Google Scholar] [CrossRef]
- Potts, P.P.; Kalaee, A.A.S.; Wacker, A. A thermodynamically consistent Markovian master equation beyond the secular approximation. New Journal of Physics 2021, 23, 123013. [Google Scholar] [CrossRef]
- Dhar, A.; Saito, K.; Hänggi, P. Nonequilibrium density-matrix description of steady-state quantum transport. Phys. Rev. E 2012, 85, 011126. [Google Scholar] [CrossRef] [PubMed]
- Mahan, G.D. Many-Particle Physics, 2000. ed., 3rd ed.; Springer US: Boston, MA, 2000. [Google Scholar]
- Kumar, P.; Webb, C.M.; Stafford, C.A. Work Sum Rule for Open Quantum Systems, 2024. arxiv:2402.18855.
1 |
throughout this paper. |
2 |
S is diagonal in this basis. |
3 |
We define a partially pure state to be any state of the system containing definite occupancies, . |
|
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. |
© 2024 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).