1. Introduction
In the Bardeen, Cooper, and Schrieffer (BCS) model [
1], the Hamiltonian
of electrons in a superconducting metal contains interaction terms proportional to the operators
, where
is a pair annihilation operators, and
annihilates a single particle Fermionic state having momentum
and spin state
. The operator
can be expressed as
, where
is the expectation value of
in thermal equilibrium, and
. In the mean field approximation (MFA), the term
is disregarded [see Equation (18.307) of Ref. [
2]]. This approximation leads to a mean field Hamiltonian
, which can be analytically diagonalized by implementing a Bogoliubov transformation.
The MFA greatly simplifies the many-body problem under study; however, it yields some predictions that are arguably inconsistent with what is expected from the original Hamiltonian
. Particle number is conserved by
, and consequently, it is expected that in steady state,
. In contrast,
, which is proportional to the BCS energy gap, can become finite in the MFA. Moreover, the ground state of the mean-field Hamiltonian
is continuously degenerate, whereas the ground state of the BCS Hamiltonian
is generically non-degenerate. The question of MFA validity is related to the spontaneous symmetry breaking in the Higgs mechanism [
3].
It was pointed out that the MFA can be, at least partially, justified in the thermodynamical limit. Particle number conservation implies that
in steady states, where
is the pair number operator. In general, the MFA allows the violation of this conservation law (i.e., it allows non-zero values of
in steady state). However, it was shown that in the MFA, both
and
are proportional to the volume of the system [
4], and thus, the violation of particle number conservation becomes insignificant in the thermodynamical limit. The mean field approach has been supported in Ref. [
5] by showing that the Ginzburg–Levanyuk parameter is typically small for electrons in metals. Moreover, it was argued in Ref. [
6] that the BCS interaction between pairs has an infinite range, and consequently exact solution of the BCS Hamiltonian
can be derived using a MFA. It was shown in Ref. [
7] that the Bogoliubov inequality, together with a variational calculation and some assumptions, can lead to the MFA Hamiltonian
. Another attempt to rigorously derive the MFA Hamiltonian
, which is based on Wick’s theorem [
8], has been presented in [
9,
10]. However, this derivation employs a relation, which can be derived from Wick’s theorem only for the case of Gaussian states [see Equation (16.131) of Ref. [
2]]. In contrast, the thermal equilibrium state that is derived from the BCS Hamiltonian
is generically non-Gaussian.
The current study is motivated by the arguably limited range of validity of the MFA, and by the difficulty in reconciling between the spontaneous symmetry breaking occurring in the superconducting state, and particle number conservation [
11,
12,
13]. An alternative approach, which is based on a recently proposed hypothesis that disentanglement spontaneously occurs in quantum systems, is explored. As is shown below, the conjecture that disentanglement plays a role in superconductivity is falsifiable, since it yields predictions that are distinguishable from what is derived from MFA-based models. In the current study, the Fermi–Hubbard model [
14,
15,
16,
17,
18,
19,
20,
21,
22,
23] is employed to study the effect of disentanglement on both superconducting order parameter and current-phase relation (CPR) of a weak link [
24].
2. Disentanglement
According to the spontaneous disentanglement hypothesis, time evolution for the reduced density operator
is governed by a modified master equation given by [
25,
26,
27,
28,
29]
where
ℏ is the Planck’s constant,
is the Hamiltonian, the operator
is allowed to depend on
, and
. The operator
is given by
, where both rates
and
are positive, and both operators
and
are Hermitian. The operator
, which gives rise to thermalization [
30,
31], is given by
, where
is the Helmholtz free energy operator [
32],
is the thermal energy inverse,
is the Boltzmann’s constant, and
T is the temperature.
For the case of a system composed of indistinguishable particles, the disentanglement operator
is derived from two–particle interaction (TPI) [
33]. The term in the Hamiltonian
accounting for TPI is denoted by
. In a basis that diagonalizes the TPI, the operator
is expressed in terms of the operators
, where
is a number operator associated with the
j’th single-particle state. In that basis, each term in
proportional to
contributes to
, a term proportional to
, where
. The term
gives rise to suppression of
, with a rate proportional to
, where the covariance
is defined by
[see Equation (
1)]. Alternatively, the covariance
can be expressed as
, where
is the probability that state
j is occupied, and
is the probability that states
and
are both occupied.
3. Fermi–Hubbard Model
Consider an array of sites occupied by Fermions. Single-site occupation energy, nearest neighbors hopping, and TPI are characterized by the real parameters
,
t, and
U, respectively. The creation and annihilation operators corresponding to site
l with spin state
are denoted by
and
, respectively. The operators
and
satisfy Fermionic anti-commutation relations. The Fermi–Hubbard Hamiltonian
is given by
, where the single-particle part
is
where
denotes that
and
are nearest neighbors, the TPI part is given by
and the Fermionic number operator
is given by
.
The term
in the TPI part
[see Equation (
3)] can be expressed as
, where
. In the MFA, i.e., when the term
is disregarded, it is well known that the Fermi–Hubbard model supports a superconducting phase for particular realizations [
34].
As was discussed above, disentanglement gives rise to the suppression of the covariance
. In the rapid disentanglement approximation [
35], it is assumed that the rate of disentanglement
is sufficiently large to allow disregarding the term
. In this limit, the disentanglement-based model yields predictions that are identical to what is derived from the standard (i.e., without disentanglement) Fermi–Hubbard model, when the MFA is implemented, and thus, the disentanglement-based model in this limit can account for superconductivity, in the same way that the mean field Fermi–Hubbard model can.
In the current study, the effect of disentanglement is explored, without assuming that
is sufficiently large to validate the rapid disentanglement approximation. As is demonstrated below, for some cases, analytical results can be derived from the modified master Equation (
1), provided that the size of the under study system is kept sufficiently small. However, since the rapid disentanglement approximation is not implemented, analysis commonly becomes intractable in the macroscopic limit.
For the relatively simple systems to be discussed below, it is assumed that the Fermi–Hubbard array is one-dimensional; the number of sites, which is denoted by
L, is finite; and the array has a ring configuration; thus, the last (
) hopping term
[see Equation (
2)] is taken to be given by
.
4. Truncation Approximation
For some cases, dynamics governed by the modified master Equation (
1) can be simplified by implementing a truncation approximation. In this approximation, the operators
and
are replaced by
and
, respectively, where
P is a projection operator. For a two-level truncation approximation, the projection
is expressed as
, where
and
are two orthonormal state vectors (i.e.,
and
). The density operator
for that case is expressed in terms of the real vector
as
where
is the Pauli matrix vector. Similarly, the Hamiltonian is expressed as
, where
is real. It is assumed that
, where
, and both the number
and the vector
are real.
The entropy operator
can be expressed as
, where
and
, and the operator
as
, where
,
, and
[recall the identity
, and note that the Pauli matrices are all trace-less]. The modified master Equation (
1) yields an equation of motion for
, given by
Note that, generally, depends on , and that the vector is orthogonal to , provided that (i.e., represents a pure state, for which ).
When the Hamiltonian
is time-independent, steady-state solutions of the modified master Equation (
1) occur at extremum points of an effective free energy
, which is given by
. In the truncation approximation,
, where
and
. For a constant
, the Helmholtz free energy
is minimized at the thermal equilibrium point
, where the unit vector
is given by
[note that
].
For the under-study Fermi–Hubbard model, and for the case of a two-site array (i.e.,
) and
, a two-level truncation approximation, which is based on a projection onto the subspace spanned by the floor
(i.e., ground) and ceiling
energy eigenstates, becomes applicable, provided that
[
33]. For the case
, the floor
and ceiling
states are given by
and
, where
,
,
, and
denotes a normalized state, where
,
,
and
. Note that the disentanglement expectation value
with respect to the state
, where the angle
is real, is given by
. Hence, in the limit
, for which
and
, the combined state
is nearly fully disentangled.
The relations
,
and
, where
, enables an analytical evaluation of the effective free energy
. The result reveals that in the low-temperature limit, and for
, a symmetry-breaking quantum phase transition occurs for this case at
. The dependency on the ratio
of steady-state values of (a) the normalized energy expectation value
and (b) purity
is shown in
Figure 1. The steady-state values are calculated by numerically integrating the modified master Equation (
1) (without employing the truncation approximation). The plot in
Figure 1b reveals that the purity
drops below unity above the phase transition occurring at
.
5. Order Parameter
The plot in
Figure 2 demonstrates the time evolution of the vector
for the case
[the truncation approximation is not being employed for the numerical integration of the modified master Equation (
1)]. The vector operator
is given by
, where
, and where
. The following holds
,
,
and
, where
and where
, and thus
(note that
). The variable
represents an order parameter.
In the low-temperature limit, and in the absence of disentanglement (i.e., for
), the ground state density operator
is a steady-state solution of the modified master Equation (
1). Note that
for the ground state
. Above the disentanglement-induced quantum phase transition, i.e., for
, the ground state becomes unstable. For the assumed parameters’ values used to generate the plot in
Figure 2, the ratio
is 50 (see figure caption). The plot shows time evolution for 16 different initial pure states, denoted by
, where
is given by
, where
[i.e.,
]. Time evolution, which is obtained by numerically integrating the modified master Equation (
1), is shown for 16 equally spaced values for the angle
in the range
. The plot demonstrates that the steady-state value of
(labelled in
Figure 2 by red × symbols) that is obtained with the initial state
is parallel to the unit vector
. Thus, for this one-dimensional model, a disentanglement-induced spontaneous symmetry breaking, which occurs for
, gives rise to finite values of the order parameter
.
6. CPR
For the case where the one-dimensional array is occupied by
spinless Fermions, the Hamiltonian
is expressed as
The Fermionic creation and annihilation operators corresponding to site are denoted by and , respectively, and the operator is given by and . It is assumed that and (i.e., all nearest neighbor site pairs, except for the pair , share the same coefficients, and ). The single-site occupation energy , hopping amplitudes t and , the phases , and the pairing amplitudes g and are all real constants. For the case of an opened chain, and , whereas and for the case of a closed ring.
The term
can be expressed as
, where
. In the MFA, for which the term
is disregarded, the resultant Hamiltonian, which is denoted by
, describes a Kitaev one-dimensional array [
36]. Note that the total number of particles is conserved by
[see Equation (
7)], whereas only the total number mod 2 is conserved by
. In the analysis below, the MFA, which generally enables violation of number conservation, is not implemented.
Consider the case where a magnetic flux given by
is externally applied to the ring’s hole, where
is real and
is the flux quantum (Planck’s constant, vacuum speed of light, and electronic charge are denoted by
h,
c, and
e, respectively). The effect of the applied flux is taken into account by setting the phases
in the Hamiltonian (7) according to
for
and
[
37,
38]. The circulating current
is calculated using the relation
[see Equation (18.142) of Ref. [
2]], where the steady-state energy expectation value
is evaluated by numerically integrating the modified master Equation (
1). For the current case, the disentanglement operator
is given by
, where
(note that
and
, where
).
The effect of disentanglement on CPR is demonstrated by the plots shown in
Figure 3. The assumed rate of disentanglement
for the plots in (a) and (b) is
, and
, respectively. For comparison, the plot in
Figure 3c displays the Beenakker–VanHouten CPR
[
39,
40], which was calculated for a single short channel of transmission
, and which is given by
, where
denotes the critical current, and [see Equation (A4) of Ref. [
41]]
The most pronounced effect of disentanglement on the CPR are the sharp features seen in
Figure 3a,b near half-integer values of the normalized applied flux
. These features do not violate the symmetry relation
, where
n is an integer. Note that some unexplained features obeying the same symmetry are visible in some spectral measurements of Josephson devices (e.g., see Figures 2 and 4 of [
42],
Figure 2 of [
43], and
Figure 2 of [
44]). Further study is needed to determine whether disentanglement can account for such experimentally observed features. Note that a variety of unconventional mechanisms, including topological and multi-band superconductivity, can give rise to CPR having features that resemble what is seen in
Figure 3a,b (e.g., see Ref. [
45]).
7. Effective Free Energy
Disentanglement is explored below by evaluating the effective free energy
for the spinless one-dimensional array in an open chain configuration. The energy eigenvalues
of
[see Equation (
7)] are shown as a function of
in
Figure 4a, for the case where
,
,
and
. For
, where
[see the black dashed vertical line in
Figure 4a], the ground state is the one-particle state
[see the blue line in
Figure 4a], whereas the two-particle state
[see the red line in
Figure 4a] becomes the ground state for
.
Consider a reduced-density operator
having matrix representation in the basis
given by
, where
is real. The truncated density operator
can be used for approximately calculating the effective free energy
for
. The dependency of
on
and
for the value
[see the green dashed vertical line in
Figure 4a] is shown in
Figure 4b (note that
does not depend on
and on
in the truncation approximation). The color-coded plot of
reveals a disentanglement-induced transition from monostability to bistability. In the low-temperature limit, and in the absence of disentanglement [i.e., in the limit
], the effective free energy
is minimized for the two-particle state
. However, for
, the system becomes bistable [see
Figure 4b].