This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Generalized Josephson effect in an asymmetric double-well potential
at finite temperatures

Kateryna Korshynska Department of Physics, Taras Shevchenko National University of Kyiv, 64/13, Volodymyrska Street, Kyiv 01601, Ukraine Fundamentale Physik für Metrologie FPM, Physikalisch-Technische Bundesanstalt PTB, Bundesallee 100, 38116 Braunschweig, Germany    Sebastian Ulbricht Fundamentale Physik für Metrologie FPM, Physikalisch-Technische Bundesanstalt PTB, Bundesallee 100, 38116 Braunschweig, Germany Institut für Mathematische Physik, Technische Universität Braunschweig, Mendelssohnstraße 3, 38106 Braunschweig, Germany
Abstract

We investigate a non-interacting many-particle bosonic system, placed in an asymmetric double-well potential. We first consider the dynamics of a single particle and determine its time-dependent probabilities to be in the left or the right well of the potential. These probabilities obey the standard Josephson equations, which in their many-particle interpretation also describe a globally coherent system, such as a Bose-Einstein condensate. This system exhibits the widely studied Josephson oscillations of the population imbalance between the wells. In our study we go beyond the regime of global coherence by developing a formalism based on an effective density matrix. This formalism gives rise to a generalization of Josephson equations, which differ from the standard ones by an additional parameter, that has the meaning of the degree of fragmentation. We first consider the solution of the generalized Josephson equations in the particular case of thermal equilibrium at finite temperatures, and extend our discussion to the non-equilibrium regime afterwards. Our model leads to a constraint on the maximum amplitude of Josephson oscillations for a given temperature and the total number of particles. A detailed analysis of this constraint for typical experimental scenarios is given.

I Introduction

In this work we study many-particle bosonic systems, as they are investigated in cold atom physics with the help of modern cooling and trapping techniques [1]. At low temperatures bosons in a weakly interacting atomic gas can be regarded as wave packets. When approaching critical temperature, those wave packets begin to overlap until they form a single matter wave at T=0T=0 [2]. This produces a unique state of matter known as Bose-Einstein condensate (BEC) [3]. Since the first experimental realization of a BEC [4, 5], its properties have been intensively studied in theory and experiment [6, 7, 2, 8, 9, 10, 11, 12] and led to various scientific applications [13, 6, 14, 15, 16, 17, 18].

The quantum nature of a BEC manifests itself, when the gas is placed in a double-well potential. Due to the coherence between the bosons in both wells, such a system exhibits a unique interference phenomenon, known as the Josephson effect [19, 7, 20, 21, 22, 23, 24, 25, 26, 27, 28]. This effect gives rise to collective oscillations of the particles, implying a time-dependent population imbalance between the wells. This behaviour has been widely studied [20, 21, 22, 23, 24, 25, 26, 27, 28] and confirmed by the first realization of a single bosonic Josephson junction in 2005 [29].

Due to the numerous difficulties in realizing ultra-cold temperatures, nowadays BECs are accessible almost exclusively for laboratory experiments in fundamental physics [6]. However, modern presicion metrology can also be performed with just laser cooled atoms, not being in BEC state, since coherence is not always an essential requirement in this area [6]. This in particular makes cooled atom-based precision tools accessible for applications in technology. The aim of this work, therefore, is to study both the ultra-cold BEC and the finite temperature regime apart from the BEC state.

In the ultra-cold BEC regime, oscillations in a double-well are theoretically described by Josephson equations, which are derived under the idealized assumption of zero temperature T=0T=0. Thus, in the present study we want to answer a natural question: how are these equations generalized for non-zero temperatures T>0T>0 and which additional effects do they predict?

The paper is organized as follows. In Sec. II we introduce the general features of a symmetric and an asymmetric double-well and define the relevant properties of its two lowest single-particle states. Based on this we derive the standard Josephson equations, which can be applied to describe a many-particle system in a globally coherent state. Then in Sec. III the single-particle states are used to build up the basis for the density matrix of a bosonic ensemble. From that in Sec. III.1 we obtain an effective density matrix, that can be used to calculate the expectation values of one-particle operators. With the help of this matrix we then derive a generalization of Josephson equations and analyze the additional physical effects they imply in Sec. III.2. While the generalized Josephson equations give rise to various modifications of the standard Josephson effect, in Sec. IV we focus on a system in thermal equilibrium and the occurrence of Josephson oscillations between the wells in the non-equilibrium regime. The conclusions are made in Sec. V.

II A single quantum particle in a double-well potential

In this paper we aim to consider a non-interacting many-particle bosonic system, which is a reasonable approximation of weakly interacting bosonic gases in many experimental situations [30, 7, 8, 10, 11, 9]. This allows us to start our discussion at the level of single-particle states in this section and to construct from that a density matrix for the many-particle system afterwards.

II.1 The symmetric double-well

In what follows we want to keep our considerations as general as possible, making them applicable to a wide range of potentials that can be realized in experiment. For convenience we will restrict our analysis to one spatial dimension, assuming that the confinement in the other two directions is sufficiently strong, such that excitations of perpendicular modes are highly suppressed (see Appendix A).

We assume a symmetric, time-independent potential V(x)=V(x)V(x)=V(-x) without further specification of its explicit shape. This potential enters the single-particle Hamiltonian

H^0=22mx2+V(x).\hat{H}_{0}=-\frac{\hbar^{2}}{2m}\partial_{x}^{2}+V(x)\,. (1)

The corresponding eigenvalue problem H^0ϕ(x)=Eϕ(x)\hat{H}_{0}\phi(x)=E\phi(x) can be solved in terms of orthogonal basis functions ϕn(x)\phi_{n}(x) with energies EnE_{n}. We assume that the Hamiltonian has at least two normalized eigenstates: the ground state ϕ0(x)\phi_{0}(x) and the first excited state ϕ1(x)\phi_{1}(x). Since the potential V(x)V(x) is symmetric, the ground state is necessarily symmetric as well, while the first excited state must be anti-symmetric, as visualized in Fig. 1a. From these states we want to construct proper left and right states ϕL/R(x)\phi_{L/R}(x), which describe the quantum particle being in the left or right part of the potential. To do this we linearly combine the energy eigenstates in the form

ϕL(x)\displaystyle\phi_{L}(x) =\displaystyle= cosξϕ0(x)+sinξϕ1(x)\displaystyle\cos\xi\,\phi_{0}(x)+\sin\xi\,\phi_{1}(x) (2)
ϕR(x)\displaystyle\phi_{R}(x) =\displaystyle= sinξϕ0(x)cosξϕ1(x)\displaystyle\sin\xi\,\phi_{0}(x)-\cos\xi\,\phi_{1}(x)\,

with a parameter ξ\xi that needs to be determined hereafter. To investigate these states, we now introduce the left- and right-side scalar products

|L+|R=|,\langle\,\cdot\,|\,\cdot\,\rangle_{L}+\langle\,\cdot\,|\,\cdot\,\rangle_{R}=\langle\,\cdot\,|\,\cdot\,\rangle\,, (3)

which imply an integral over all negative or positive xx, respectively, such that their sum gives the standard scalar product. Calculating the left- and right-side scalar products of the left and right states (2) we obtain

ϕR|ϕRR=ϕL|ϕLL\displaystyle\langle\phi_{R}|\phi_{R}\rangle_{R}=\langle\phi_{L}|\phi_{L}\rangle_{L} =\displaystyle= 1/2+sin(2ξ)ϕ0|ϕ1L\displaystyle 1/2+\sin(2\xi)\langle\phi_{0}|\phi_{1}\rangle_{L}
ϕR|ϕRL=ϕL|ϕLR\displaystyle\langle\phi_{R}|\phi_{R}\rangle_{L}=\langle\phi_{L}|\phi_{L}\rangle_{R} =\displaystyle= 1/2sin(2ξ)ϕ0|ϕ1L.\displaystyle 1/2-\sin(2\xi)\langle\phi_{0}|\phi_{1}\rangle_{L}\,. (4)

In this calculation we use ϕ0|ϕ0L/R=ϕ1|ϕ1L/R=1/2\langle\phi_{0}|\phi_{0}\rangle_{L/R}=\langle\phi_{1}|\phi_{1}\rangle_{L/R}=1/2, which is clear by symmetry considerations. The scalar product ϕ0|ϕ1L=ϕ0|ϕ1R\langle\phi_{0}|\phi_{1}\rangle_{L}=-\langle\phi_{0}|\phi_{1}\rangle_{R} can be chosen real and positive by adapting the phases of the energy eigenstates. Recalling that the scalar products (4) represent the probabilities to find the particles of state ϕL/R(x)\phi_{L/R}(x) in the left or right part of the potential, we can maximize this probability by the choice ξ=π/4\xi=\pi/4, leading to the states

ϕL(x)\displaystyle\phi_{L}(x) =\displaystyle= 12[ϕ0(x)+ϕ1(x)]\displaystyle\frac{1}{\sqrt{2}}[\phi_{0}(x)+\phi_{1}(x)]
ϕR(x)\displaystyle\phi_{R}(x) =\displaystyle= 12[ϕ0(x)ϕ1(x)]\displaystyle\frac{1}{\sqrt{2}}[\phi_{0}(x)-\phi_{1}(x)]\, (5)

as the best fitting left and right states in the potential V(x)V(x), shown in Fig. 2a. The explicit shape of this potential now only enters our consideration via the scalar product ϕ0|ϕ1L\langle\phi_{0}|\phi_{1}\rangle_{L}, which is a measure of how good the states ϕL/R(x)\phi_{L/R}(x) actually represent a particle at the left or right side of the potential. In particular, we will call all such potentials a symmetric double-well, in which ϕ0|ϕ1L=1/2ϵ\langle\phi_{0}|\phi_{1}\rangle_{L}=1/2-\epsilon with a sufficiently small ϵ>0\epsilon>0, such that

ϕR|ϕRR=ϕL|ϕLL\displaystyle\langle\phi_{R}|\phi_{R}\rangle_{R}=\langle\phi_{L}|\phi_{L}\rangle_{L} =\displaystyle= 1ϵ\displaystyle 1-\epsilon
ϕR|ϕRL=ϕL|ϕLR\displaystyle\langle\phi_{R}|\phi_{R}\rangle_{L}=\langle\phi_{L}|\phi_{L}\rangle_{R} =\displaystyle= ϵ,\displaystyle\epsilon\,, (6)

and the ϕL/R(x)\phi_{L/R}(x) appropriately describe a single quantum particle in a left or right well state.

Refer to caption
Figure 1: Example geometries of (a) a symmetric double-well potential V(x)V(x) with the two lowest energy eigenstates ϕ0/1(x)\phi_{0/1}(x) of a contained quantum particle, and (b) an asymmetric double-well V(x)+δV(x)V(x)+\delta V(x) with corresponding states ψ0/1(x)\psi_{0/1}(x).

In the next section, we will use these definitions for the symmetric double-well to extend our analysis to a double-well potential with an additional potential step.

II.2 The asymmetric double-well

To describe a particle in an asymmetric double-well potential we extend the Hamiltonian

H^=H^0+δV(x)\hat{H}=\hat{H}_{0}+\delta V(x) (7)

by a small potential step

δV(x)={V0x<00elsewhere.\delta V(x)=\left\{\begin{array}[]{ll}V_{0}&\quad x<0\\ 0&\quad\mathrm{elsewhere}\,\,.\end{array}\right.\, (8)

The solutions of the corresponding eigenvalue problem

H^ψn(x)=E~nψn(x)\hat{H}\psi_{n}(x)=\tilde{E}_{n}\psi_{n}(x) (9)

can be obtained in the framework of first order perturbation theory, up to linear order in V0/EV_{0}/E, where E=(E0+E1)/2E=(E_{0}+E_{1})/2. For that we use the states from Sec. II.1 and treat δV(x)\delta V(x) as a small perturbation of the symmetric double-well.

Refer to caption
Figure 2: Left and right well states in (a) a symmetric and (b) an asymmetric double-well potential. The functional form of both coincides, regardless of the value of V0EV_{0}\ll E.

It is well known that in perturbation theory, states strongly affect each other when they have similar energies. For the symmetric double-well (6) the separation of the ground and excited state from other possible states is in the order of EE. In contrast, their own energies are much closer to each other, i.e., E1E0=ΔEEE_{1}-E_{0}=\Delta E\ll E. This is due to the fact that the scalar product ϕ0|ϕ1L=1/2ϵ\langle\phi_{0}|\phi_{1}\rangle_{L}=1/2-\epsilon only slightly deviates from 1/21/2, such that the states ϕ0(x)ϕ1(x)\phi_{0}(x)\approx\phi_{1}(x) closely resemble each other for x<0x<0, as also can be seen in Fig. 1a. In consequence, acting on them with the Hamiltonian H^1\hat{H}_{1} gives similar energy eigenvalues E0E1E_{0}\approx E_{1}, which become degenerated in the ultimate case ϵ=0\epsilon=0. We, thus, can restrict our investigation to the corrections of ϕ0(x)\phi_{0}(x) by ϕ1(x)\phi_{1}(x) and vise versa, while we have to deal with the perturbation theory for nearly degenerated states [31]. For that we make the ansatz ψn(x)=cn0ϕ0(x)+cn1ϕ1(x){\psi}_{n}(x)=c^{0}_{n}\phi_{0}(x)+c^{1}_{n}\phi_{1}(x) in Eq. (9) and integrate this equation with the states ϕ0(x)\phi^{*}_{0}(x) and ϕ1(x)\phi^{*}_{1}(x), respectively. This, together with the normalization of the corrected states, yields six equations for the six constants cn0c^{0}_{n}, cn1c^{1}_{n}, and E~n\tilde{E}_{n} with n=0,1n=0,1. The only non-trivial, linearly independent solution is given by

ψ0(x)\displaystyle\psi_{0}(x) =\displaystyle= 𝒞(ϕ0(x)V0ΔE+ΔE2+V02ϕ1(x))\displaystyle\mathcal{C}\left(\phi_{0}(x)-\frac{V_{0}}{\Delta E+\sqrt{\Delta E^{2}+V_{0}^{2}}}\,\phi_{1}(x)\right) (10)
ψ1(x)\displaystyle\psi_{1}(x) =\displaystyle= 𝒞(ϕ1(x)+V0ΔE+ΔE2+V02ϕ0(x))\displaystyle\mathcal{C}\left(\phi_{1}(x)+\frac{V_{0}}{\Delta E+\sqrt{\Delta E^{2}+V_{0}^{2}}}\,\phi_{0}(x)\right)\,

with 𝒞=12(1+ΔE/ΔE2+V02)1/2\mathcal{C}=\frac{1}{\sqrt{2}}\left(1+\Delta E/\sqrt{\Delta E^{2}+V_{0}^{2}}\right)^{1/2}. Those states satisfy Eq. (9) to linear order in V0/EV_{0}/E for the corrected energies

E~0/1=E+V0212ΔE2+V02,\tilde{E}_{0/1}=E+\frac{V_{0}}{2}\mp\frac{1}{2}\sqrt{\Delta E^{2}+V_{0}^{2}}\,, (11)

where we assume that ϵV0/E\epsilon V_{0}/E can be neglected, cf. Eq. (6). The ψ0/1(x)\psi_{0/1}(x), hence, are energy eigenstates of the asymmetric double-well, illustrated in Fig. 1b.

As for the symmetric double-well in the last section, we now can define the left and right well states ψL/R(x)\psi_{L/R}(x) in the case of the asymmetric double-well potential

ψL(x)\displaystyle\psi_{L}(x) =\displaystyle= cosξψ0(x)+sinξψ1(x)\displaystyle\cos\xi\,\psi_{0}(x)+\sin\xi\,\psi_{1}(x)
ψR(x)\displaystyle\psi_{R}(x) =\displaystyle= sinξψ0(x)cosξψ1(x)\displaystyle\sin\xi\,\psi_{0}(x)-\cos\xi\,\psi_{1}(x)\, (12)

in total analogy to Eq. (2), but this time using the energy eigenstates ψ0/1(x)\psi_{0/1}(x) as a basis.

We now again ask for the optimal states that maximize ψR|ψRR\langle\psi_{R}|\psi_{R}\rangle_{R} and ψL|ψLL\langle\psi_{L}|\psi_{L}\rangle_{L}. Inserting Eq. (10) into Eq. (12) we can formulate this question in the basis ϕ0/1(x)\phi_{0/1}(x) of the symmetric double-well problem. In this basis the optimal states ϕL/R(x)\phi_{L/R}(x) are already known and given by Eq. (5). Therefore, we find that those states are optimal for both, the symmetric and the asymmetric case, i.e, ψL(x)=ϕL(x)\psi_{L}(x)=\phi_{L}(x) and ψR(x)=ϕR(x)\psi_{R}(x)=\phi_{R}(x), see Fig. 2. In Eq. (12) this is achieved by the choice

ξ=arcsin[12(1+V0/ΔE2+V02)1/2],\xi=\arcsin\left[\frac{1}{\sqrt{2}}\left(1+V_{0}/\sqrt{\Delta E^{2}+V_{0}^{2}}\right)^{1/2}\right]\,, (13)

which for V0=0V_{0}=0 recovers the symmetric case ξ=π/4\xi=\pi/4.

Having found the optimal left and right states for the asymmetric double-well potential, we can investigate their dynamics, which is governed by the Hamiltonian operator (7). With ψL/R(x)\psi_{L/R}(x) being a superposition of energy eigenstates, it is clear that they do not satisfy an eigenvalue equation for a particular energy. Instead, we obtain the coupled equations

H^ψL(x)\displaystyle\hat{H}\psi_{L}(x) =\displaystyle= ELψL(x)+KψR(x)\displaystyle E_{L}\psi_{L}(x)+K\psi_{R}(x)
H^ψR(x)\displaystyle\hat{H}\psi_{R}(x) =\displaystyle= ERψR(x)+KψL(x),\displaystyle E_{R}\psi_{R}(x)+K\psi_{L}(x)\,, (14)

with the newly introduced constants

EL=E+V0,ER=E,K=ΔE2.E_{L}=E+V_{0}\,,\quad E_{R}=E\,,\quad K=-\frac{\Delta E}{2}\,. (15)

With that we find a coupling ΔE\sim\Delta E between the left and right state, which vanishes when the energies E0/1E_{0/1} degenerate, as it is the case when the wells are seperated by a very high potential barrier. Moreover, the difference ELER=V0E_{L}-E_{R}=V_{0} coincides with the potential step (8), which is a well known feature of a Josephson junction [32].

II.3 Josephson equations and many-particle interpretation

In the last section we analyzed the eigenvalue equation H^ψ0/1=E~0/1ψ0/1\hat{H}\psi_{0/1}=\tilde{E}_{0/1}\psi_{0/1}, holding for the time-dependent, but stationary states ψ0/1(x,t)=exp(iE~0/1t/)ψ0/1(x)\psi_{0/1}(x,t)=\exp(-i\tilde{E}_{0/1}t/\hbar)\psi_{0/1}(x). From the spatial wave functions ψ0(x)\psi_{0}(x) and ψ1(x)\psi_{1}(x) the left and right states ψL(x)\psi_{L}(x) and ψR(x)\psi_{R}(x) were constructed, cf. Eqs. (12). Now we want to investigate the evolution of the states ψL(x,t)\psi_{L}(x,t) and ψR(x,t)\psi_{R}(x,t), which are not stationary states of a particular energy EL/RE_{L/R}, solely, as becomes clear in Eqs. (14). Instead, we can assume a time-dependent superposition of ψL(x)\psi_{L}(x) and ψR(x)\psi_{R}(x), giving rise to a single wave-function

ψ(x,t)=wL(t)ψL(x)+wR(t)ψR(x).\psi(x,t)=w_{L}(t)\psi_{L}(x)+w_{R}(t)\psi_{R}(x)\,. (16)

By using the action (14) of the one-particle Hamiltonian H^\hat{H} on the left and right well states, as well as the orthogonality relation ψL|ψR=0\langle\psi_{L}|\psi_{R}\rangle=0, we obtain

iw˙L(t)\displaystyle i\hbar\dot{w}_{L}(t) =\displaystyle= ELwL(t)+KwR(t)\displaystyle E_{L}w_{L}(t)+Kw_{R}(t)
iw˙R(t)\displaystyle i\hbar\dot{w}_{R}(t) =\displaystyle= ERwR(t)+KwL(t).\displaystyle E_{R}w_{R}(t)+Kw_{L}(t).

These equations describe the evolution of the coefficients wL/R(t)w_{L/R}(t), where |wL/R(t)|2|w_{L/R}(t)|^{2} represent the time-dependent probabilities to find the particle in the left or right well, respectively. Here, the normalization condition would be given by |wL(t)|2+|wR(t)|2=1|w_{L}(t)|^{2}+|w_{R}(t)|^{2}=1.

While, so far we have applied the wave function (16) to describe a single particle, it can also be used to represent a many-particle system, as long as that system behaves coherently, i.e, is in a pure state. In this case, we only need to renormalize Eq. (16) to obtain |wL(t)|2+|wR(t)|2=N|w_{L}(t)|^{2}+|w_{R}(t)|^{2}=N, where NN is the total number of bosons. Then, the |wL/R(t)|2=NL/R(t)|w_{L/R}(t)|^{2}=N_{L/R}(t) represent the population of the left and right well, respectively.

Staying with the many-particle interpretation and by defining wL/R(t)=NL/R(t)eiθL/R(t)w_{L/R}(t)=\sqrt{N_{L/R}(t)}e^{i\theta_{L/R}(t)} we find the coupled differential equations

Z˙\displaystyle\hbar\dot{Z} =\displaystyle= 2K1Z2sinθ\displaystyle 2K\sqrt{1-Z^{2}}\sin\theta (17)
θ˙\displaystyle\hbar\dot{\theta} =\displaystyle= ELER2KZcosθ1Z2,\displaystyle E_{L}-E_{R}-\frac{2KZ\cos\theta}{\sqrt{1-Z^{2}}}, (18)

for the fractional population imbalance Z(t)=(NL(t)NR(t))/NZ(t)=(N_{L}(t)-N_{R}(t))/N and the phase difference θ(t)=θR(t)θL(t)\theta(t)=\theta_{R}(t)-\theta_{L}(t). These are standard Josephson equations, as they are used in literature to describe the standard Josephson effect [20, 21, 22, 23, 24, 25, 26]. As we recapitulate in Appendix B, the general solution of these equations leads to oscillations of the observable population imbalance Z(t)Z(t) with the frequency ΔE2+V02/\sqrt{\Delta E^{2}+V_{0}^{2}}/\hbar.

Having found these familiar expressions for a many-particle system in a pure state, in what follows we will investigate what happens when the pure state assumption is not initially made. We will show, that this leads to a generalization of Josephson equations, opening up a way to investigate many-particle dynamics in the double-well beyond the regime of global coherence.

III Many-particle bosonic system in a double-well

In the last section we have studied the dynamics of a single particle based on the generic features of a double-well potential giving rise to specific eigenstates and energies. The following ideas, however, can be applied to an arbitrary two-state system.

Having defined single-particle states |ψ0/1|\psi_{0/1}\rangle in a given geometry, we can elaborate a description of a many-particle system. For that we assume the Hamiltonian of a bosonic ensemble of NN non-interacting particles, given by

^=i=1NH^(xi)\hat{\mathcal{H}}=\sum_{i=1}^{N}\hat{H}(x_{i})\, (19)

with the single-particle Hamiltonian (7). The assumption of non-interacting particles is valid, as long as the one-dimensional particle density is much smaller than the inverse of the scattering length for the considered bosons [33]. The consequences of this assumption for the thermalization process are discussed in Appendix A. In general, one would describe such a system by introducing the hermitian density operator [34]

ρ^=N1=0NN~1=0NpN1N~1(t)|ΨN1ΨN~1|,\hat{\rho}=\sum_{N_{1}=0}^{N}\sum_{\tilde{N}_{1}=0}^{N}p_{N_{1}\tilde{N}_{1}}(t)|\Psi_{N_{1}}\rangle\langle\Psi_{\tilde{N}_{1}}|\,, (20)

where pN1N1p_{N_{1}N_{1}} is the probability that the system is in a state with N1N_{1} excited particles. The normalization condition for the density matrix reads N1=0NpN1N1=1\sum_{N_{1}=0}^{N}p_{N_{1}N_{1}}=1. Moreover, the state |ΨN1|\Psi_{N_{1}}\rangle in Eq. (20) is given by

|ΨN1=1N!(NN1)!N1!𝑑x1𝑑xN×j(N1)ψσjN1(1)(x1)ψσjN1(2)(x2)ψσjN1(N)(xN)|x1,,xN,|\Psi_{N_{1}}\rangle=\frac{1}{\sqrt{N!(N-N_{1})!N_{1}!}}\int dx_{1}...dx_{N}\\ \times\sum_{j(N_{1})}\psi_{\sigma_{j}^{N_{1}}(1)}(x_{1})\psi_{\sigma_{j}^{N_{1}}(2)}(x_{2})...\psi_{\sigma_{j}^{N_{1}}(N)}(x_{N}){\color[rgb]{0,0,0}|x_{1},\dots,x_{N}\rangle}, (21)

where the sum is over all possible configurations j(N1)j(N_{1}) of single-particle states for a fixed number N1N_{1}. Each configuration is labeled by an index jj and characterised by a vector σj=(σj(1),σj(2),,σj(N))\mathbf{\sigma}_{j}=(\sigma_{j}(1),\sigma_{j}(2),...,\sigma_{j}(N)), defining the state of each particle.

Since the state |ΨN1|\Psi_{N_{1}}\rangle is constructed from the single-particle states ψ0(x)\psi_{0}(x) and ψ1(x)\psi_{1}(x), the σj(i)\sigma_{j}(i) can only take the values 0 or 11. However, the particle ii is still allowed to be in a superposition of both states, since the coefficients pN1,N~1p_{N_{1},\tilde{N}_{1}} for N1N~1N_{1}\neq\tilde{N}_{1} are not necessarily zero.

III.1 Effective density matrix description

In this section we will reformulate the description of a many-particle system by the density matrix (20) in terms of an effective density matrix in the basis of the one-particle states |ψ0|\psi_{0}\rangle and |ψ1|\psi_{1}\rangle, which appear as the two lowest states in the double-well potential. In the following we are particularly interested in the occupation probabilities of these states, needed to investigate the Josephson effect.

For this purpose, we consider a generic single-particle operator

𝒪^=O^𝟙𝟙N-times+𝟙O^𝟙+ +𝟙𝟙O^\hat{\mathcal{O}}=\underbrace{\hat{O}\oplus\mathbb{1}\oplus\dots\oplus\mathbb{1}}_{N\textnormal{-times}}+\mathbb{1}\oplus\hat{O}\oplus\dots\oplus\mathbb{1}+\dots{\\[-8.5359pt] }+\mathbb{1}\oplus\mathbb{1}\oplus\dots\oplus\hat{O}

in the NN-particle Hilbert space, constructed from the operator O^\hat{O}, acting on a single-particle state. For instance, the occupation number operator N^0/1\hat{N}_{0/1} is given by the particular choice O^=|ψ0/1ψ0/1|\hat{O}=|\psi_{0/1}\rangle\langle\psi_{0/1}| in the formula above. The expectation value of a many-particle operator then reads

𝒪^ρ=tr(ρ^𝒪^)=N1=0NN~1=0NpN1N~1ψN~1|𝒪^|ψN1=N1=0NpN1N1[N1ψ1|O^|ψ1+(NN1)ψ0|O^|ψ0]+N1=0N1(N1+1)(NN1)[pN1+1,N1ψ0|O^|ψ1+pN1,N1+1ψ1|O^|ψ0],\langle\hat{\mathcal{O}}\rangle_{\rho}=\mathrm{tr}(\hat{\rho}\hat{\mathcal{O}})=\sum_{N_{1}=0}^{N}\sum_{\tilde{N}_{1}=0}^{N}p_{N_{1}\tilde{N}_{1}}\langle\psi_{\tilde{N}_{1}}|\hat{\mathcal{O}}|\psi_{N_{1}}\rangle\\ =\sum_{N_{1}=0}^{N}p_{N_{1}N_{1}}[N_{1}\langle\psi_{1}|\hat{O}|\psi_{1}\rangle+(N-N_{1})\langle\psi_{0}|\hat{O}|\psi_{0}\rangle]\\ +\sum_{N_{1}=0}^{N-1}\sqrt{(N_{1}+1)(N-N_{1})}[p_{N_{1}+1,N_{1}}\langle\psi_{0}|\hat{O}|\psi_{1}\rangle\\ +p_{N_{1},N_{1}+1}\langle\psi_{1}|\hat{O}|\psi_{0}\rangle], (22)

cf. [35]. In front of the expectation values ψi|O^|ψj\langle\psi_{i}|\hat{O}|\psi_{j}\rangle we now can read off the coefficients αij\alpha_{ij}, which contain all information accessible by the measurement of one-particle observables:

α00\displaystyle\alpha_{00} =\displaystyle= N1=0N(NN1)pN1N1\displaystyle\sum_{N_{1}=0}^{N}(N-N_{1})p_{N_{1}N_{1}}
α11\displaystyle\alpha_{11} =\displaystyle= N1=0NN1pN1N1\displaystyle\sum_{N_{1}=0}^{N}N_{1}p_{N_{1}N_{1}}
α01\displaystyle\alpha_{01} =\displaystyle= N1=0N1(N1+1)(NN1)pN1+1,N1\displaystyle\sum_{N_{1}=0}^{N-1}\sqrt{(N_{1}+1)(N-N_{1})}p_{N_{1}+1,N_{1}}
α10\displaystyle\alpha_{10} =\displaystyle= α01.\displaystyle\alpha^{*}_{01}.

We now formally rearrange Eq. (22) and introduce the effective density matrix ρ^e\hat{\rho}_{\mathrm{e}}:

𝒪^ρ=[ψ0|ψ1|][α00α01α10α11]O^[|ψ0|ψ1]=tr(ρ^eO^).\langle\hat{\mathcal{O}}\rangle_{\rho}=\begin{bmatrix}\langle\psi_{0}|&\langle\psi_{1}|\\ \end{bmatrix}\begin{bmatrix}\alpha_{00}&\alpha_{01}\\ \alpha_{10}&\alpha_{11}\\ \end{bmatrix}\hat{O}\begin{bmatrix}|\psi_{0}\rangle\\ |\psi_{1}\rangle\\ \end{bmatrix}=\mathrm{tr}(\hat{\rho}_{\mathrm{e}}\hat{O})\,. (23)

As we show in Appendix C, this effective density matrix, which we obtained from the calculation of one-particle expectation values, coincides with the reduced density matrix of a single particle in a bath of all the other N1N-1 bosons [36]. It therefore can be interpreted as a description of the many-particle ensemble by an average boson.

Further, we are interested in the population of the left and right potential wells rather than the occupation of the global ground and excited states |ψ0/1|\psi_{0/1}\rangle. Therefore, in the following we will describe the system in the basis of left and right well states |ψL/R|\psi_{L/R}\rangle, introduced in Sec. II. Here this is done by the matrix transformation

[|ψL|ψR]=[cosξ|ψ0+sinξ|ψ1sinξ|ψ0cosξ|ψ1]=T^[|ψ0|ψ1],\begin{bmatrix}|\psi_{L}\rangle\\ |\psi_{R}\rangle\\ \end{bmatrix}=\begin{bmatrix}\cos\xi|\psi_{0}\rangle+\sin\xi|\psi_{1}\rangle\\ \sin\xi|\psi_{0}\rangle-\cos\xi|\psi_{1}\rangle\\ \end{bmatrix}=\hat{T}\begin{bmatrix}|\psi_{0}\rangle\\ |\psi_{1}\rangle\\ \end{bmatrix}\,, (24)

where the parameter ξ\xi is given by Eq. (13) for the asymmetric double-well potential. The matrix T^\hat{T} now can be used to express the effective density matrix ρ^eLR=T^ρ^eT^1\hat{\rho}_{eLR}=\hat{T}\hat{\rho}_{e}\hat{T}^{-1} in the left and right well basis. In general, this hermitian 2×22\times 2 matrix can be parameterized by

ρ^eLR=[NL(t)A(t)eiθ(t)A(t)eiθ(t)NR(t)],\hat{\rho}_{\mathrm{e}LR}=\begin{bmatrix}N_{L}(t)&A(t)e^{i\theta(t)}\\ A(t)e^{-i\theta(t)}&N_{R}(t)\\ \end{bmatrix}, (25)

where NL(t)N_{L}(t) and NR(t)N_{R}(t) are the occupation numbers of the left and right well, respectively. Moreover, by tr(ρ^eLR)=N=NL(t)+NR(t)\mathrm{tr}(\hat{\rho}_{\mathrm{e}LR})=N=N_{L}(t)+N_{R}(t) the total number of particles is conserved. The non-diagonal complex matrix elements describe the interference between left and right well states and therefore induce the coupling between the wells by a mixing parameter A(t)A(t) and the phase difference θ(t)\theta(t) between them.

In what follows, we will derive the equations of motion for ρ^eLR\hat{\rho}_{\mathrm{e}LR}, which will generalize the standard Josephson equations (17) and (18) to the case where the system is not described by a single wave-function.

III.2 Generalized Josephson equations

With the help of the effective density matrix (25), we can go beyond the description of a many-particle system by a pure state. This will give us the opportunity to consider a wider range of physical setups. For instance, this will enable us to describe bosonic systems at finite temperatures.

The evolution of the effective density matrix is given by the Liouville equation

itρ^eLR=[^eLR,ρ^eLR],i\hbar\frac{\partial}{\partial t}\hat{\rho}_{\mathrm{e}LR}=[\hat{\mathcal{H}}_{\mathrm{e}LR},\hat{\rho}_{\mathrm{e}LR}], (26)

with the Hamiltonian operator ^eLR=H^|ψLψL|+H^|ψRψR|\hat{\mathcal{H}}_{\mathrm{e}LR}=\hat{H}|\psi_{L}\rangle\langle\psi_{L}|+\hat{H}|\psi_{R}\rangle\langle\psi_{R}|, where the single-particle Hamiltonian H^\hat{H} is given by Eq. (7). By using the action of H^\hat{H} on the left and right well states |ψL/R|\psi_{L/R}\rangle given in Eq. (14), the Liouville equation reduces to the three coupled differential equations

Z˙\displaystyle\hbar\dot{Z} =\displaystyle= 4KNAsinθ\displaystyle 4\frac{K}{N}A\sin\theta (27)
θ˙\displaystyle\hbar\dot{\theta} =\displaystyle= ELERKNZAcosθ\displaystyle E_{L}-E_{R}-\frac{KNZ}{A}\cos\theta (28)
A˙A\displaystyle\frac{\dot{A}}{A} =\displaystyle= [θ˙+EREL]tanθ.\displaystyle\left[\dot{\theta}+\frac{E_{R}-E_{L}}{\hbar}\right]\tan\theta\,. (29)

In what follows we want to draw attention to the additional physical effects that are described by the generalized Josephson equations (27) - (29) in comparison to the standard ones (17),(18). For this purpose, we use the fact that these equations can be rewritten in the form

Z˙\displaystyle\hbar\dot{Z} =\displaystyle= 2Kf2Z2sinθ\displaystyle 2K\sqrt{f^{2}-Z^{2}}\sin\theta (30)
θ˙\displaystyle\hbar\dot{\theta} =\displaystyle= ELER2KZcosθf2Z2,\displaystyle E_{L}-E_{R}-\frac{2KZ\cos\theta}{\sqrt{f^{2}-Z^{2}}}, (31)

as we show in Appendix D. The structure of these equations is obtained by formally integrating Eq. (29) for A(t)A(t) and eliminating the mixing parameter in the Josephson equations (27) and (28) for Z(t)Z(t) and θ(t)\theta(t). The representation (30),(31) of the generalized Josephson equations only deviates from the standard ones by the newly introduced constant parameter ff. At the level of the effective density matrix, ff is fixed by the initial condition ρ^eLR(t=0)\hat{\rho}_{\mathrm{e}LR}(t=0).

For f=1f=1 we recover the standard Josephson equations, i.e., dynamics in the pure state regime. At the level of the effective density matrix this corresponds to the special choice A(t)=NL(t)NR(t)=N1Z(t)2/2A(t)=\sqrt{N_{L}(t)N_{R}(t)}=N\sqrt{1-Z(t)^{2}}/2. In this case the effective density matrix is a projector |ψψ||\psi\rangle\langle\psi| to one of its eigenvectors, representing a globally coherent state, such that 𝒪^ρ=ψ|𝒪^|ψ\langle\hat{\mathcal{O}}\rangle_{\rho}=\langle\psi|\hat{\mathcal{O}}|\psi\rangle. In this regime the system can be described by the single wave function ψ\psi, given by Eq. (16).

For general values of ff the eigenvalues of the effective density matrix ρ^e\hat{\rho}_{\mathrm{e}} are given by N(1±f)/2N(1\pm f)/2, see Appendix D. Recalling the interpretation of the effective density matrix as the reduced density matrix for an average boson, these eigenvalues have to be non-negative, to ensure non-negative probabilities. This leads to the restriction |f|1|f|\leq 1. Following the discussion of BEC fragmentation, given in [37, 38], we choose f>0f>0 and use the term degree of fragmentation for the 1f1-f parameter, implying f=1f=1 for a non-fragmented globally coherent state and f=0f=0 for two incoherent fragmented states. The ff parameter can be associated with the commonly used degree of coherence, or coherence factor, which defines the visibility of interference fringes in an interference experiment with fragmented BECs [3]. However, notice that the same terms are used differently in the studies of quantum fluctuations in BECs [39, 40], where a strongly interacting system is discussed (see also Appendix D).

IV Generalized Josephson effect

Let us now briefly summarize what we have done so far. We started with a non-interacting NN-particle bosonic system, described by the (N+1)×(N+1)(N+1)\times(N+1) dimensional density matrix (20). Then, we formulated an effective density matrix (25), which contains all the information about the many-particle system relevant for the calculation of one-particle observables. In this framework we derived the generelized Josephson equations (27)-(29) and their general solution for the evolution of the effective density matrix elements, see Appendix E. This allows us to describe various regimes of many-particle dynamics: in what follows, we first use the generalized Josephson equations to investigate a bosonic system in thermal equilibrium. The thermalization process, which leads to this state, is not part of the present discussion. However, we briefly elaborate on that in Appendix A.

In particular, we will utilize the description of the many-particle system by a canonical ensemble to determine the parameters of the effective density matrix in terms of the temperature TT, the particle number NN and the energy difference ΔE~=E~1E~0\Delta\tilde{E}=\tilde{E}_{1}-\tilde{E}_{0}. Afterwards, we extend the discussion to the case of non-equilibrium leading to oscillatory dynamics. These oscillations are studied in the pure state and in the generalized cases. Moreover, the corresponding behaviour of the degree of fragmentation 1f1-f is analyzed. This will allow us to investigate how the interplay of temperature, particle number, and the geometry of the double-well potential affects the particle dynamics.

IV.1 Effective density matrix in thermal equilibrium

In the following, we want to describe a closed many-particle bosonic system in thermal equilibrium of finite temperature TT. In thermodynamics and statistical physics this is typically done by assuming a canonical ensemble with the density matrix

ρ^=𝒵1eβH^,\hat{\rho}=\mathcal{Z}^{-1}e^{-\beta\hat{H}}\,, (32)

where we introduced β=1/kBT\beta=1/k_{\mathrm{B}}T, and H^\hat{H} is given by Eq. (19) [34]. Moreover, the partition function 𝒵=tr(eβH^)\mathcal{Z}=\mathrm{tr}(e^{-\beta\hat{H}}) ensures the normalization trρ^=1\mathrm{tr}\hat{\rho}=1.

Projecting the density operator (32) to the basis of states |ΨN1|\Psi_{N_{1}}\rangle, as done in Eq. (20), we can read off the density matrix elements

pN1N~1\displaystyle p_{N_{1}\tilde{N}_{1}} =\displaystyle= ΨN1|𝒵1eβH^|ΨN~1\displaystyle\langle\Psi_{N_{1}}|\mathcal{Z}^{-1}e^{-\beta\hat{H}}|\Psi_{\tilde{N}_{1}}\rangle
=\displaystyle= 𝒵1eβ(NN1)E~0βN1E~1δN1,N1~.\displaystyle\mathcal{Z}^{-1}e^{-\beta(N-N_{1}){\tilde{E}}_{0}-\beta N_{1}{\tilde{E}}_{1}}\delta_{N_{1},\tilde{N_{1}}}\,.

We find that the original density matrix ρ^\hat{\rho} is diagonal, which also results in a diagonal effective density matrix ρ^e\hat{\rho}_{\mathrm{e}} in the |ψ0/1|\psi_{0/1}\rangle basis with

α00\displaystyle\alpha_{00} =\displaystyle= 𝒵1N1=0N(NN1)eβE~1N1βE~0(NN1)\displaystyle\mathcal{Z}^{-1}\sum_{N_{1}=0}^{N}(N-N_{1})e^{-\beta{\tilde{E}}_{1}N_{1}-\beta{\tilde{E}}_{0}(N-N_{1})} (33)
=\displaystyle= 1+Ne(N+1)βΔE~(N+1)eNβΔE~(1eβΔE~)(e(N+1)βΔE~1)\displaystyle\frac{1+Ne^{(N+1)\beta\Delta{\tilde{E}}}-(N+1)e^{N\beta\Delta{\tilde{E}}}}{(1-e^{-\beta\Delta{\tilde{E}}})(e^{(N+1)\beta\Delta{\tilde{E}}}-1)}
α11\displaystyle\alpha_{11} =\displaystyle= 𝒵1N1=0NN1eβE~1N1βE~0(NN1)\displaystyle\mathcal{Z}^{-1}\sum_{N_{1}=0}^{N}N_{1}e^{-\beta{\tilde{E}}_{1}N_{1}-\beta{\tilde{E}}_{0}(N-N_{1})} (34)
=\displaystyle= e(N+1)βΔE~+N(N+1)eβΔE~(eβΔE~1)(e(N+1)βΔE~1)\displaystyle\frac{e^{(N+1)\beta\Delta{\tilde{E}}}+N-(N+1)e^{\beta\Delta{\tilde{E}}}}{(e^{\beta\Delta{\tilde{E}}}-1)(e^{(N+1)\beta\Delta{\tilde{E}}}-1)}
α01\displaystyle\alpha_{01} =\displaystyle= α10=0,\displaystyle\alpha_{10}=0, (35)

where ΔE~=ΔE2+V02\Delta{\tilde{E}}={\color[rgb]{0,0,0}\sqrt{\Delta E^{2}+V_{0}^{2}}}, cf. Eq. (11), and consistently α00+α11=N\alpha_{00}+\alpha_{11}=N. This gives rise to the effective density matrix

ρ^e=[α0000α11]\hat{\rho}_{\mathrm{e}}=\begin{bmatrix}\alpha_{00}&0\\ 0&\alpha_{11}\\ \end{bmatrix} (36)

as the starting point for our investigation of the generalized Josephson effect in thermal equilibrium.

IV.2 Static solution in thermal equilibrium

In this work we assume a Bose gas with negligibly small interactions. In such a system the establishment of thermal equilibrium takes a very long time Δtth/ΔE\Delta t_{\mathrm{th}}\gg\hbar/\Delta E that is much larger than the typical time scale of dynamics in the double-well potential (see also Appendix A). In this section, we assume that the system has already undergone this thermalization process and reached thermal equilibrium. Then the elements of the effective density matrix ρ^e\hat{\rho}_{\mathrm{e}} are given by Eqs. (33)-(35) and we are ready to apply our formalism to describe the generalized Josephson effect in this regime.

For that purpose we consider the effective density matrix in the left and right well basis. Applying the transformation T^\hat{T} from Eq. (24) to Eq. (36), we obtain

ρ^eLR(V0)=N2[1001]+δN012ΔE2+V02[V0ΔEΔEV0],{\color[rgb]{0,0,0}\hat{\rho}_{\mathrm{e}LR}(V_{0})=\frac{N}{2}\begin{bmatrix}1&0\\ 0&1\\ \end{bmatrix}+\frac{\delta N_{01}}{2\sqrt{\Delta E^{2}+V_{0}^{2}}}\begin{bmatrix}-V_{0}&\Delta E\\ \Delta E&V_{0}\\ \end{bmatrix},} (37)

where we introduced the population imbalance δN01(N,βΔE~)=α00α11\delta N_{01}(N,\beta\Delta{\tilde{E}})=\alpha_{00}-\alpha_{11} between ground and excited state. Using Eq. (37) as a specific initial condition for these solutions we find

Z(t)=V0ΔE~δN01N,A(t)=ΔEΔE~δN012,θ(t)=0,{\color[rgb]{0,0,0}Z(t)=-\frac{V_{0}}{\Delta\tilde{E}}\frac{\delta N_{01}}{N}\,\,,\quad A(t)=\frac{\Delta E}{\Delta\tilde{E}}\frac{\delta N_{01}}{2}\,\,,\quad\theta(t)=0\,,} (38)

such that no oscillations between the wells occur. In consequence, Eq. (37) is not only the initial condition but a static solution for the effective density matrix, i.e., for the dynamics of the system. We recognize that Eqs. (38) represent the unique solution of the generalized Josephson equations for NN particles in a given double-well geometry at temperature TT.

Refer to caption
Figure 3: Degree of condensation f=δN01/Nf=\delta N_{01}/N as a function of temperature TT in thermal equilibrium. Above the populations of the ground E~0{\tilde{E}}_{0} and excited states E~1{\tilde{E}}_{1} are shown schematically.

From the solution (37) we find, that in thermal equilibrium, the degree of fragmentation 1f=1δN01/N1-f=1-\delta N_{01}/N equals the population imbalance between the excited state |ψ1|\psi_{1}\rangle and the ground state |ψ0|\psi_{0}\rangle. In this case ff reaches from f=0f=0 for kBTΔE~k_{\mathrm{B}}T\gg\Delta{\tilde{E}} to f=1f=1 for kBTΔE~k_{\mathrm{B}}T\ll\Delta{\tilde{E}}, as shown in Fig. 3. The latter particularly holds true for T=0T=0, where α00=N\alpha_{00}=N, such that the bosons form a global BEC in the double-well system. Hence, in thermal equilibrium f[0,1]f\in[0,1] can be interpreted as the degree of condensation. In the literature such a parameter is also used to introduce T0T\neq 0 effects in the Gross-Pitaevskii equation phenomenologically [41]. However, in our model it arises from first principles of statistics of a many-particle quantum system and has a wider interpretation in the non-equilibrium case.

We can summarize the findings of this section with the evident statement that in the case of thermal equilibrium the solutions of the generalized Josephson equations are static and the system exhibits no Josephson oscillations. However, the insights from this section will be of great value also for the study of the non-equilibrium case in the next section.

IV.3 Josephson effect in non-equilibrium regime

We now will discuss how Josephson oscillations appear as an out-of-equilibrium phenomenon in a particular experimental scenario [42]. First a cloud of atoms is thermalized to the equilibrium configuration with some population imbalance between the wells, induced by the energy difference ELER=V0iE_{L}-E_{R}=V_{0}^{i}. Then, the potential step between the wells is instantly lowered to a new value V0f<V0iV_{0}^{f}<V_{0}^{i}, bringing the system out of equilibrium. Directly after that, the system evolves according to Eqs. (27)-(29), but with ELER=V0fE_{L}-E_{R}=V_{0}^{f}. The initial conditions for the system are given by ρ^eLR(V0i)\hat{\rho}_{\mathrm{e}LR}(V_{0}^{i}) from Eq. (37). Plugging that into the general solutions to the generalized Josephson equations we present in Appendix E, we obtain the population imbalance

Z(t)=V0fΔE2+(V0i)2δN01iNV0iV0fΔE2+(V0i)2δN01iNcos(ΔEt),Z(t)=-\frac{V_{0}^{f}}{\sqrt{\Delta E^{2}+(V_{0}^{i})^{2}}}\frac{\delta N_{01}^{i}}{N}\\ -\frac{V_{0}^{i}-V_{0}^{f}}{\sqrt{\Delta E^{2}+(V_{0}^{i})^{2}}}\frac{\delta N_{01}^{i}}{N}\cos\left(\frac{\Delta Et}{\hbar}\right), (39)

showing an oscillatory behavior. Here, for the sake of simplicity we only consider terms to linear order in the parameter V0f/ΔEV_{0}^{f}/\Delta E, which is assumed to be small in our further discussion. This includes the case considered in Ref. [42], where the final double-well is symmetric V0f=0V_{0}^{f}=0.

As can be seen in Eq. (39), the amplitude of oscillations is proportional to the difference between the initial and final potential step V0iV0fV_{0}^{i}-V_{0}^{f}, which also quantifies how far the system is from the initial equilibrium state. In the limit V0iΔEV_{0}^{i}\gg\Delta E we observe the maximum possible amplitude of oscillations δN01i/N\delta N_{01}^{i}/N for given NN, ΔE\Delta E and TT, see Fig. 4 for particular temperatures. In the opposite case if V0iΔEV_{0}^{i}\ll\Delta E the amplitude of Josephson oscillations is proportional to a small ratio (V0iV0f)/ΔE(V_{0}^{i}-V_{0}^{f})/\Delta E.

The ff parameter in the non-equilibrium regime reads f=δN01i/Nf=\delta N_{01}^{i}/N, where δN01i=α00iα11i\delta N_{01}^{i}=\alpha_{00}^{i}-\alpha_{11}^{i} is the initial population imbalance between the ground and first excited state with the energies E~0i\tilde{E}_{0}^{i} and E~1i\tilde{E}_{1}^{i}, respectively. Moreover, the ff parameter remains independent of the new potential step V0fV_{0}^{f} and is preserved from the initial thermal equilibrium condition, discussed in Sec. IV.2. In general the values V0iV_{0}^{i}, V0fV_{0}^{f} and ΔE\Delta E are known from the experimental setup [42], which allows us to deduce ff from the oscillation amplitude. The ff parameter depends only on the number of bosons NN and ΔE~i/(kBT)\Delta\tilde{E}^{i}/(k_{B}T). Thus, by the measurement of the amplitude of Josephson oscillations in the Bose gas it is possible to determine the thermodynamic temperature it had before the potential step was lowered V0iV0fV_{0}^{i}\to V_{0}^{f}.

The pure state case f=1f=1, for which standard Josephson equations hold, is realized for kBTΔE~ik_{B}T\ll\Delta\tilde{E}^{i}, allowing for the maximum possible oscillation amplitude. The opposite regime of f=0f=0 can be achieved either in the case of high temperatures or in the case of nearly degenerated initial energy levels E~0/1i\tilde{E}_{0/1}^{i} in the double-well, i.e, ΔE~i0\Delta\tilde{E}^{i}\approx 0. Nearly degenerated energy levels appear for very high potential barriers between the two wells. In such a geometry the populations of each well become weakly coupled, as can be seen in the Eqs. (14) and (15). This is exactly the regime, where the system can be seen as two separate, weakly interacting systems (often referred to as coupled BECs), as it is done, e.g., in the Bose-Hubbard model [3, 23].

Refer to caption
Figure 4: Oscillations of the population imbalance Z(t)Z(t) as a function of dimensionless time ΔEt/\Delta Et/\hbar. The initial potential step is chosen as V0i=ΔEV_{0}^{i}=\Delta E and the final potential step V0f=0V_{0}^{f}=0. The number of particles is fixed to N=103N=10^{3}, while the temperature is T=2.5×104T=2.5\times 10^{-4} K (red dotted), T=106T=10^{-6} K (blue dashed), and T=108T=10^{-8} K (blue solid).

The equilibrium position of the oscillations (39) is given by the averages Z(t)t=V0f/ΔE2+(V0i)2×δN01i/N\langle Z(t)\rangle_{t}=-V_{0}^{f}/\sqrt{\Delta E^{2}+(V_{0}^{i})^{2}}\times\delta N_{01}^{i}/N and θ(t)t=0\langle\theta(t)\rangle_{t}=0 over one period of oscillations. We see that this equilibrium position depends on the initial value of potential step V0iV_{0}^{i}, implying that the system ‘remembers’ the initial condition. By direct comparison with the Eqs. (38) we find that this equilibrium position does not represent a static thermal equilibrium solution of the generalized Josephson equations for the new double-well with the potential step V0fV_{0}^{f}. Thus, to reach the new static equilibrium the system must ‘forget’ its initial condition during the process of thermalization. Therefore, this process must not only cause a damping of the oscillations but also has to shift the equilibrium position to the position of a, yet undefined, new static thermal equilibrium. While this process itself is not accessible within our model directly (see also Appendix A), we can deduce some conclusions about its final state. The exact nature of the thermalization process defines to which final thermal equilibrium state the system tends to. All these possible final thermal equilibrium states are labeled by their degree of fragmentation 1f1-f. For each of those f[0,1]f\in[0,1] there is a unique static solution of the generalized Josephson equations (30) - (31) and the degree of fragmentation for the final equilibrium state does not necessarily coincides with the initial one. Assuming a fixed degree of fragmentation 1f=1δN01i/N1-f=1-\delta N_{01}^{i}/N, inherited from the initial condition, one would need to allow for a loss of energy. Otherwise, assuming the energy to be conserved, one has to account for an increase of the degree of fragmentation to allow for thermalization to happen. This means that the system is less coherent after an adiabatic thermalization process.

A more detailed analysis of the thermalization mechanism in dependence of the underlying physical processes can be an interesting future perspective of our study.

IV.4 Implications for cold atom experiments

In what follows, we consider an NN-particle system of initial temperature TT and discuss the implications of our model for different experimental setups. In Table 1 we give the upper limit for the amplitude of Josephson oscillations (39), determined by the initial degree of condensation f=δN01i/Nf=\delta N_{01}^{i}/N for experimentally relevant scenarios, related to typical temperatures and cooling techniques [43]. The corresponding oscillations for N=103N=10^{3} and different temperatures are visualized in Fig. 4.

On its way to ultra-cold temperatures the many-particle bosonic system undergoes different cooling stages of characteristic temperatures. The first stage is the formation of a longitudinal supersonic beam with a narrow velocity distribution (about T8T\propto 8 K) [44].

If lower temperatures are pursued, the beam can be placed in a magneto-optical trap (MOT) and can be laser-cooled to temperatures of T2.5×104T\propto 2.5\times 10^{-4} K [45]. In this regime the ff parameter sensitively depends on the particle number, ranging from N103106N\propto 10^{3}-10^{6} in typical experiments [7, 18, 17, 21, 29, 23, 42]. While for 10310^{3} particles we see that the amplitude of Josephson oscillations is restricted to maximally 0.5%0.5\% of all particles, for 10610^{6} this amplitude is barely restricted.

At the next cooling stage the atoms are placed in an optical dipole trap reaching typical temperatures in the range of T106T\propto 10^{-6} K [46]. For low particle numbers 103\propto 10^{3}, in this regime, it is necessary to use not the standard, but the generalized Josephson equations with a degree of fragmentation of 1f0.251-f\approx 0.25 (see Table 1). To describe such a cold (but not yet ultra-cold) Bose gas correctly is of interest for modern precision metrology [6].

Finally, evaporative cooling and adiabatic expansion can lead to the BEC formation at T108T\propto 10^{-8} K [4] with almost all particles condensed in the ground state, such that δN01iN\delta N^{i}_{01}\approx N. In this regime the system is almost fully coherent and can be represented by a single wave-function to good approximation. Here, the standard Josephson equations are sufficient to describe the dynamics of the many-particle system. In the BEC regime our model puts no significant restriction on the amplitude of Josephson oscillations, in agreement with the experimental observations [7, 17, 29, 23, 42].

Table 1: Limit on maximum amplitude of Josephson oscillations δN01i/N\delta N_{01}^{i}/N for typical temperatures TT and particle numbers NN, related to specific experimental scenarios. An exemplary oscillation frequency of ΔE/=103rad/s\Delta E/\hbar=10^{3}\,\mathrm{rad/s} out of the typical range of 102rad/s104rad/s10^{2}\,\mathrm{rad/s}-10^{4}\,\mathrm{rad/s} [7, 17, 29, 23, 42].
Temperature regime NN f=δN01i/Nf=\delta N_{01}^{i}/N
Optical molasses 10310^{3} 0.0051
or MOT 10410^{4} 0.0508
T=2.5×104T=2.5\times 10^{-4} K 10510^{5} 0.4443
10610^{6} 0.9345
Collimated beam 10310^{3} 0.7401
(transverse) 10410^{4} 0.9739
T=106T=10^{-6} K 10510^{5} 0.9974
10610^{6} 0.9997
10310^{3} 0.9983
BEC T=108T=10^{-8} K 10410^{4} 0.9998
10510^{5} 0.9999
10610^{6} 1.0000

V Conclusions

In this paper we derived a generalization of the standard Josephson equations, which can be used to describe bosonic many-particle systems in the non-coherent regime, apart from Bose-Einstein condensation. In particular, we apply this formalism to study a system of NN quantum particles in an asymmetric double-well potential at finite temperatures. For this purpose we first construct an effective density matrix, which allows us to calculate the expectation values of one-particle operators. For this density matrix, we derive the generalized Josephson equations as the central part of our theory. These equations define the evolution of the population imbalance, the phase difference and a newly introduced mixing parameter, which together provide a full description of the bosonic system on the level of one-particle operators. The newly introduced mixing parameter allows us to investigate the many-particle bosonic system, that not necessarily has to be in the BEC phase.

The generalization of Josephson equations leads to additional physical effects. To analyze these effects we introduce an additional parameter f1f\leq 1 with 1f1-f having a meaning of the degree of fragmentation. The ultimate case f=1f=1 corresponds to the pure state of the system described by the standard Josephson equations, as they are derived in literature [7, 20, 21, 22, 23].

The approach presented in this paper does not require any restrictions on the initial many-particle density matrix, up to its defining properties, implying f[0,1]f\in[0,1]. Hence, it is suitable to describe a wide range of physical scenarios beyond the pure state case. In thermal equilibrium of finite temperature T>0T>0 the approach yields a static solution with constant population imbalance and zero phase difference between the wells. In this regime the parameter f=δN01/Nf=\delta N_{01}/N equals the fractional population imbalance between the ground and excited energy eigenstates and, hence, has the meaning of the degree of condensation.

To discuss the non-equilibrium regime we considered an initially thermalized Bose gas in an asymmetric double-well potential. Then, in the modeled experimental scenario [42], the potential step between the wells is instantly lowered. This leads to oscillatory dynamic in a double-well system. We found that the oscillation amplitude depends on temperature TT, total number of particles NN, energy difference ΔE\Delta E, as well as the initial and final potential steps V0iV_{0}^{i} and V0fV_{0}^{f}. In experiment one can access the values of V0iV_{0}^{i}, V0fV_{0}^{f}, ΔE\Delta E and NN, implying that the knowledge of the amplitude of the Josephson oscillations may allow to determine the temperature of the system. This opens up an intriguing possibility of quantum thermometry.

We found that for a fixed double-well geometry the amplitude is limited by the ff parameter, which coincides with the initial degree of condensation δN01i/N\delta N_{01}^{i}/N of the system before the potential step was lowered. This restriction becomes recognizable for cold (but not ultra-cold) Bose gases at temperatures T106T\geq 10^{-6} K and vanishes in the BEC regime of T108T\propto 10^{-8} K. This analysis highlights that the Josephson effect as a quantum interference phenomenon is more pronounced for ultra-cold Bose gases [3]. However, the presented results complement this well-known statement by a quantitative discussion of the suppression of Josephson oscillations in the finite temperature regime. While this conclusion fits the intuition about the Josephson effect, that is expected to vanish at higher temperatures, we want to point out, that this results sensitively depend on the experimental setup, i.e., on how non-equilibrium is obtained.

The matter of this paper is the generalization of the Josephson equations based on the statistical properties of the quantum system, which allows to go beyond the BEC regime. However, the applicability of our model is limited by the assumption that the interactions between the bosons are negligibly small, which leads to a long time of thermalization in comparison to oscillation time scales. Moreover, we assumed that the trapped Bose is strongly confined in all but one direction, the total number of bosons remains constant and the system is closed, i.e., does not interact with the environment. For instance, by including interaction between the bosons, one will have an additional energy scale in the model, while implying higher-order correlation between the bosons and allowing for thermalization in a three-dimensional analysis. These may lead to a significant modification of the model defining the possible directions in which the present study could be extended.

Acknowledgements.
The authors want to thank Dorothee Tell, Yuriy M. Bidasyuk, Claus Lämmerzahl, and Andrey Surzhykov for comments and valuable discussions. Funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy—EXC 2123 QuantumFrontiers—390837967.

Appendix A Thermalization in a bosonic gas

In this Appendix we briefly address the constraints on our model under which the Bose gas can be assumed to be an effective 1D system of non-interacting particles. We address how such a system can reach thermal equilibrium and state the conditions under which thermalization processes can be neglected on the time scales of Josephson oscillations.

The effective 1D model, considered in this paper, holds for Bose gases which are strongly confined in the other two directions [33]. This confinement translates into large energies ωz=ωy=ωΔE\hbar\omega_{z}=\hbar\omega_{y}=\hbar\omega_{\perp}\gg\Delta E, needed to excite a perpendicular state. In order to suppress this excitations in a system of finite temperature TT, the inequality ω>kBT\hbar\omega_{\perp}>k_{B}T must hold. In this case, the wavefunction can be decoupled into a product of a transverse ground state and a 1D state ψ(x,t)\psi(x,t) we are interested in.

For establishment of thermal equilibrium one would need to have a redistribution of energy leading to thermalization. However, in a 1D collision of two identical bosons no energy exchange and therefore no thermalization would happen. This redistribution can be achieved only by including 3D inter-particle collisions, which would populate the transverse excited states, and which are neglected in the Hamiltonian (19). These collisional interactions can be present but negligible, as long as the linear particle density n1Dn_{1D} and the scattering length asa_{s} of the bosons satisfy n1Das1n_{1D}a_{s}\ll 1 [33]. In this case the time scale of thermalization Δtth\Delta t_{th} is much larger than the period h/ΔEh/\Delta E of Josephson oscillations. This is in agreement with the analysis in Sec. III and Sec. IV. In particular, within this approximation, we neglect damping due to reestablishment of thermal equilibrium in Sec. IV.3.

For a detailed study on the thermalization process, see Ref. [33].

Appendix B Standard Josephson solutions

In this Appendix the technical details of solving the pure state Josephson equations (17),(18) are discussed. To simplify the problem, we divide these equations following the rule dZdt(dθdt)1=dZdtdtdθ=dZdθ\frac{dZ}{dt}\left(\frac{d\theta}{dt}\right)^{-1}=\frac{dZ}{dt}\frac{dt}{d\theta}=\frac{dZ}{d\theta} to obtain a differential equation for the Z(θ)Z(\theta) function. Then we introduce new functions x=cosθx=\cos\theta and Z=sinyZ=\sin y, in terms of which the differential equation reads

dxdy=xtanyELER2K.\frac{dx}{dy}=x\tan y-\frac{E_{L}-E_{R}}{2K}.

In the following we will denote the dimensionless ratio in this equation as δ=(ELER)/(2K)\delta=(E_{L}-E_{R})/(2K). In the case of the asymmetric double-well (15) we find δ=V0/ΔE\delta=-V_{0}/\Delta E. The general solution is x(y)=1Δρs2cosy+c2δcosyδtanyx(y)=\frac{\sqrt{1-\Delta\rho_{\mathrm{s}}^{2}}}{\cos y}+\frac{c_{2}\delta}{\cos y}-\delta\tan y, so cosθ=1Δρs2+c2δ1Z(t)2δZ1Z2\cos\theta=\frac{\sqrt{1-\Delta\rho_{\mathrm{s}}^{2}}+c_{2}\delta}{\sqrt{1-Z(t)^{2}}}-\delta\frac{Z}{\sqrt{1-Z^{2}}}. Substituting this result into the first equation (18), one gets:

dZ1+βZγZ2=αdt,\frac{dZ}{\sqrt{1+\beta Z-\gamma Z^{2}}}=\alpha dt\,,

where the following notations are introduced β=2δ(1Δρs2+c2δ)Δρs22c2δ1Δρs2c22δ2\beta=\frac{2\delta\left(\sqrt{1-\Delta\rho_{\mathrm{s}}^{2}}+c_{2}\delta\right)}{\Delta\rho_{\mathrm{s}}^{2}-2c_{2}\delta\sqrt{1-\Delta\rho_{\mathrm{s}}^{2}}-c_{2}^{2}\delta^{2}}, γ=1+δ2Δρs22c2δ1Δρs2c22δ2\gamma=\frac{1+\delta^{2}}{\Delta\rho_{\mathrm{s}}^{2}-2c_{2}\delta\sqrt{1-\Delta\rho_{\mathrm{s}}^{2}}-c_{2}^{2}\delta^{2}} and α=2KΔρs22c2δ1Δρs2c22δ2\alpha=\frac{2K}{\hbar}\sqrt{\Delta\rho_{\mathrm{s}}^{2}-2c_{2}\delta\sqrt{1-\Delta\rho_{\mathrm{s}}^{2}}-c_{2}^{2}\delta^{2}}. Performing the shift Z~=Zβ2γ\tilde{Z}=Z-\frac{\beta}{2\gamma} one obtains the solution for Z(t)Z(t) and θ(t)\theta(t):

Z(t)=δ(c2δ+1Δρs2)1+δ2+Δρs2+δ(δc22δ2c21Δρs2))1+δ2sin(2K(a)1+δ2t+ϕ0s+δα),\displaystyle Z(t)=\frac{\delta(c_{2}\delta+\sqrt{1-\Delta\rho_{\mathrm{s}}^{2}})}{1+\delta^{2}}+\frac{\sqrt{\Delta\rho_{\mathrm{s}}^{2}+\delta(\delta-c_{2}^{2}\delta-2c_{2}\sqrt{1-\Delta\rho_{\mathrm{s}}^{2}}))}}{1+\delta^{2}}\sin\left(\frac{2K^{\mathrm{(a)}}}{\hbar}\sqrt{1+\delta^{2}}t+\phi_{0\mathrm{s}}+\delta\alpha\right),
θ(t)=arccos(1Δρs2+c2δ1Z(t)2δZ(t)1Z(t)2).\displaystyle\theta(t)=\arccos\left(\frac{\sqrt{1-\Delta\rho_{\mathrm{s}}^{2}}+c_{2}\delta}{\sqrt{1-Z(t)^{2}}}-\delta\frac{Z(t)}{\sqrt{1-Z(t)^{2}}}\right).

To obtain the solution in the case of a symmetric double-well, one needs to set δ=0\delta=0 in the expressions above. The latter solution contains the two integration constants Δρs\Delta\rho_{\mathrm{s}} and ϕ0s\phi_{0\mathrm{s}}, while in the asymmetric case the two additional constants c2c_{2} and δα\delta\alpha appear.

To linear order in δ=V0/ΔE\delta=-V_{0}/\Delta E, i.e, for a small asymmetry of the double-well potential, we have

Z(t)=V0ΔE1Δρs2+(Δρs+c2V0ΔE1Δρs2Δρs)sinψtV0ΔEαΔρscosψt,Z(t)=-\frac{V_{0}}{\Delta E}\sqrt{1-\Delta\rho_{\mathrm{s}}^{2}}\\ +\left(\Delta\rho_{\mathrm{s}}+c_{2}\frac{V_{0}}{\Delta E}\frac{\sqrt{1-\Delta\rho_{\mathrm{s}}^{2}}}{\Delta\rho_{\mathrm{s}}}\right)\sin\psi_{\mathrm{t}}\\ -\frac{V_{0}}{\Delta E}\alpha\Delta\rho_{\mathrm{s}}\cos\psi_{\mathrm{t}},
θ(t)=arccos(1Δρs21Δρs2sin2ψt)V0ΔEΔρs3sinψtcosψtc2cosψtαΔρs21Δρs2sinψtΔρs(1Δρs2sin2ψt),\theta(t)=\arccos\left(\frac{\sqrt{1-\Delta\rho_{\mathrm{s}}^{2}}}{\sqrt{1-\Delta\rho_{\mathrm{s}}^{2}\sin^{2}\psi_{\mathrm{t}}}}\right)\\ -\frac{V_{0}}{\Delta E}\frac{\Delta\rho_{\mathrm{s}}^{3}\sin\psi_{\mathrm{t}}\cos\psi_{\mathrm{t}}-c_{2}\cos\psi_{\mathrm{t}}-\alpha\Delta\rho_{\mathrm{s}}^{2}\sqrt{1-\Delta\rho_{\mathrm{s}}^{2}}\sin\psi_{\mathrm{t}}}{\Delta\rho_{\mathrm{s}}(1-\Delta\rho_{\mathrm{s}}^{2}\sin^{2}\psi_{\mathrm{t}})},

where we introduced the notation ψt=ϕ0sΔEt/\psi_{\mathrm{t}}=\phi_{0\mathrm{s}}-\Delta Et/\hbar.

Appendix C Interpretation of effective density matrix

In this Appendix we will establish the interpretation of the effective density matrix ρ^e\hat{\rho}_{e}, which determines expectation values of one-particle operators, cf. Eq. (23). Such operators project the Hilbert space of the many-particle states |ΨN1|\Psi_{N_{1}}\rangle to a two-dimensional space of single-particle states |ψ0/1|\psi_{0/1}\rangle. Now we will prove that such projection corresponds to reducing the original density matrix ρ^=N1N~1pN1,N~1|ΨN1ΨN~1|\hat{\rho}=\sum_{N_{1}\tilde{N}_{1}}p_{N_{1},\tilde{N}_{1}}|\Psi_{N_{1}}\rangle\langle\Psi_{\tilde{N}_{1}}| to a reduced single-particle density matrix ρ^1\hat{\rho}_{1}, as used, e.g., in Ref. [36].

The system, which is the whole NN-particle ensemble, is now divided into two subsystems H^=H^1𝟙+𝟙H^B\hat{H}=\hat{H}_{1}\oplus\mathbb{1}+\mathbb{1}\oplus\hat{H}_{B}, which are a single 1st particle and a bath of all other N1N-1 particles. The whole ensemble is described by the |ΨN1|\Psi_{N_{1}}\rangle states from Eq. (21), while the energy eigenbasis of the bath reads

|Φn1=1(N1)!(N1n1)!n1!dx2dxN××j(n1)ψσjn1(2)(x2)ψσjn1(N)(xN)|x2,,xN,|\Phi_{n_{1}}\rangle=\frac{1}{\sqrt{(N-1)!(N-1-n_{1})!n_{1}!}}\int dx_{2}...dx_{N}\times\\ \times\sum_{j(n_{1})}\psi_{\sigma_{j}^{n_{1}}(2)}(x_{2})...\psi_{\sigma_{j}^{n_{1}}(N)}(x_{N}){\color[rgb]{0,0,0}|x_{2},\dots,x_{N}\rangle},

in a complete analogy with the notations in Eq. (21).

In order to extract the important information about the state of a 1st particle, one needs to trace the total density matrix ρ^\hat{\rho} over the bath states |Φn1|\Phi_{n_{1}}\rangle

ρ^1=trBρ^=n1=0N1N1=0NN~1=0NpN1N~1(t)Φn1|ΨN1BΨN~1|Φn1B,\hat{\rho}_{1}=\mathrm{tr}_{\mathrm{B}}\hat{\rho}\\ =\sum_{n_{1}=0}^{N-1}\sum_{N_{1}=0}^{N}\sum_{\tilde{N}_{1}=0}^{N}p_{N_{1}\tilde{N}_{1}}(t)\langle\Phi_{n_{1}}|\Psi_{N_{1}}\rangle_{\mathrm{B}}\langle\Psi_{\tilde{N}_{1}}|\Phi_{n_{1}}\rangle_{\mathrm{B}},

where Φn1|ΨN1B\langle\Phi_{n_{1}}|\Psi_{N_{1}}\rangle_{\mathrm{B}} is a scalar product in the bath Hilbert space only. Hence, it is an element of the single-particle Hilbert space in the |ψ0/1|\psi_{0/1}\rangle basis. Due to the orthogonality of the one-particle states, the scalar product Φn1|ΨN1B\langle\Phi_{n_{1}}|\Psi_{N_{1}}\rangle_{\mathrm{B}} does not vanish only for values of n1=N11,N1n_{1}=N_{1}-1,N_{1}. The scalar product ΨN~1|Φn1B\langle\Psi_{\tilde{N}_{1}}|\Phi_{n_{1}}\rangle_{\mathrm{B}} is non-zero only for n1=N~11,N~1n_{1}=\tilde{N}_{1}-1,\tilde{N}_{1}. Then the expression above can be simplified to just one summation over N1N_{1} with four terms, which can be calculated by using the explicit expressions of |ΨN1|\Psi_{N_{1}}\rangle and |Φn1|\Phi_{n_{1}}\rangle. For instance, the product with n1=N1n_{1}=N_{1} reads

ΦN1|ΨN1B=1N(NN1)i=1NN1|ψ0(xi)=NN1N|ψ0(x).\langle\Phi_{N_{1}}|\Psi_{N_{1}}\rangle_{\mathrm{B}}=\frac{1}{\sqrt{N(N-N_{1})}}\sum_{i=1}^{N-N_{1}}|\psi_{0}(x_{i})\rangle\\ =\sqrt{\frac{N-N_{1}}{N}}|\psi_{0}(x)\rangle\,.

The last equality in the expression above holds only effectively, i.e., for the calculation of one-particle observables O^\hat{O}, introduced in Sec. III.1. Without referring to the particular type of observables, this approximation can be treated as the mean-field approach with one averaged coordinate xx instead of many-particle coordinates {xi}\{x_{i}\}. The same can be done for the other term

ΦN11|ΨN1=1NN1i=1N1|ψ1(xi)=N1N|ψ1(x).\langle\Phi_{N_{1}-1}|\Psi_{N_{1}}\rangle=\frac{1}{\sqrt{NN_{1}}}\sum_{i=1}^{N_{1}}|\psi_{1}(x_{i})\rangle=\sqrt{\frac{N_{1}}{N}}|\psi_{1}(x)\rangle.

This simplification yields the reduced single-particle density matrix

ρ^1=1N(α00|ψ0ψ0|+α01|ψ1ψ0|+α10|ψ0ψ1|+α11|ψ1ψ1|),\hat{\rho}_{1}=\frac{1}{N}\left(\alpha_{00}|\psi_{0}\rangle\langle\psi_{0}|+\alpha_{01}|\psi_{1}\rangle\langle\psi_{0}|\right.\\ \left.+\alpha_{10}|\psi_{0}\rangle\langle\psi_{1}|+\alpha_{11}|\psi_{1}\rangle\langle\psi_{1}|\right)\,,

where the αij\alpha_{ij} coefficients are the same as defined in Sec. III.1. So we conclude that ρ^e=Nρ^1\hat{\rho}_{\mathrm{e}}=N\hat{\rho}_{1}.

Appendix D Degree of fragmentation

In the Appendix we will show how to obtain the generalized Josephson equations in the form

Z˙\displaystyle\hbar\dot{Z} =\displaystyle= 2Kf2Z2sinθ,\displaystyle 2K\sqrt{f^{2}-Z^{2}}\sin\theta, (40)
θ˙\displaystyle\hbar\dot{\theta} =\displaystyle= ELER2KZcosθf2Z2,\displaystyle E_{L}-E_{R}-\frac{2KZ\cos\theta}{\sqrt{f^{2}-Z^{2}}}, (41)

analogous to the standard Josephson equations except of f1f\neq 1. The latter equations look simpler, than their original form with the mixing parameter A(t)A(t) (27)-(29), however this way the meaning of the ff parameter is hidden. To define ff one would need to trace back to the original generalized Josephson equations and apply the initial conditions for the effective density matrix Z(t=0)Z(t=0), θ(t=0)\theta(t=0) and A(t=0)A(t=0). Now we will show how the ff parameter can be expressed via the integration constants, introduced in the Appendix above, while in the general case it reads f2=Z(t)2+(2A(t)/N)2f^{2}=Z(t)^{2}+(2A(t)/N)^{2} and represents the pure state condition if f=1f=1.

As shown in Appendix E, the general solution for the symmetric double-well reads Z(t)=Δρssin[ΔEt/+ϕ0s]Z(t)=\Delta\rho_{\mathrm{s}}\sin\left[\Delta Et/\hbar+\phi_{0\mathrm{s}}\right] and θ(t)=arctan[NeBsΔρs2cos(ΔEt/+ϕ0s)]\theta(t)=-\arctan\left[\frac{Ne^{-B_{\mathrm{s}}}\Delta\rho_{\mathrm{s}}}{2}\cos\left(\Delta Et/\hbar+\phi_{0\mathrm{s}}\right)\right] which leads to

cosθ=[1+(NeBs2)2(Δρs2Z2)]1/2.\cos\theta=\left[1+\left(\frac{Ne^{-B_{\mathrm{s}}}}{2}\right)^{2}\left(\Delta\rho_{\mathrm{s}}^{2}-Z^{2}\right)\right]^{-1/2}\,.

We substitute this expression in A(t)=eBs/cosθ(t)A(t)=e^{B_{\mathrm{s}}}/\cos\theta(t) and then plug it in the first two Josephson equations (27),( 28), which yields

Z˙\displaystyle\hbar\dot{Z} =\displaystyle= ΔE(f(s))2Z2sinθ\displaystyle-\Delta E\sqrt{(f^{\mathrm{(s)}})^{2}-Z^{2}}\sin\theta
θ˙\displaystyle\hbar\dot{\theta} =\displaystyle= ΔEZcosθ(f(s))2Z2,\displaystyle\Delta E\frac{Z\cos\theta}{\sqrt{(f^{\mathrm{(s)}})^{2}-Z^{2}}},

where (f(s))2=(2eBs/N)2+Δρs2(f^{\mathrm{(s)}})^{2}=\left(2e^{B_{\mathrm{s}}}/N\right)^{2}+\Delta\rho_{\mathrm{s}}^{2}. This immediately recovers pure state condition by taking f(s)=1f^{\mathrm{(s)}}=1.

In slightly asymmetric double-well we make an educated guess that the equations (27) - (29) can be written in a form (40), (41). Thus, Eq. (40) yields the condition

(f(a))2=Z2+(ΔE)2Z˙2sin2θ=const.(f^{\mathrm{(a)}})^{2}=Z^{2}+\left(\frac{\hbar}{\Delta E}\right)^{2}\frac{\dot{Z}^{2}}{\sin^{2}\theta}=const.

Substituting the general solution (E), (48) into the expression above, we find that indeed f(a)=constf^{\mathrm{(a)}}=const and equals

(f(a))2=(f(s))2+V0ΔEeBsN[2c1e2BsΔρssinδϕ0+N(f(s))2(δBeBsNΔρssinϕ0s)].(f^{\mathrm{(a)}})^{2}=(f^{\mathrm{(s)}})^{2}+\frac{V_{0}}{\Delta E}\frac{e^{-B_{\mathrm{s}}}}{N}\left[2c_{1}e^{2B_{\mathrm{s}}}\Delta\rho_{\mathrm{s}}\sin\delta\phi_{0}\right.\\ \left.+N(f^{\mathrm{(s)}})^{2}\left(\delta Be^{B_{\mathrm{s}}}-N\Delta\rho_{\mathrm{s}}\sin\phi_{0\mathrm{s}}\right)\right].

In the same way one can prove that the equation (41) holds for the solutions of the generalized Josephson equations.

The f=Z(t)2+(2A(t)/N)2f=\sqrt{Z(t)^{2}+(2A(t)/N)^{2}} parameter also appears when we diagonalize the effective density matrix ρ^eLR\hat{\rho}_{\mathrm{e}LR} from Eq. (25) to obtain

ρ~^e=N2[1+f001f]=N2[(1+f)|y1y1|+(1f)|y2y2|]\hat{\tilde{\rho}}_{\mathrm{e}}=\frac{N}{2}\begin{bmatrix}1+f&0\\ 0&1-f\\ \end{bmatrix}\\ =\frac{N}{2}\left[(1+f)|y_{1}\rangle\langle y_{1}|+(1-f)|y_{2}\rangle\langle y_{2}|\right] (42)

in the orthonormal basis of its eigenvectors

|y1=[1+(N2A(t)(Z(t)f))2]1/2×[|ψLNeiθ(t)2A(t)(Z(t)f)|ψR],|y_{1}\rangle=\left[1+\left(\frac{N}{2A(t)}(Z(t)-f)\right)^{2}\right]^{-1/2}\\ \times\left[|\psi_{L}\rangle-\frac{Ne^{-i\theta(t)}}{2A(t)}(Z(t)-f)|\psi_{R}\rangle\right], (43)
|y2=[1+(N2A(t)(Z(t)+f))2]1/2×[|ψLNeiθ(t)2A(t)(Z(t)+f)|ψR],|y_{2}\rangle=\left[1+\left(\frac{N}{2A(t)}(Z(t)+f)\right)^{2}\right]^{-1/2}\\ \times\left[|\psi_{L}\rangle-\frac{Ne^{-i\theta(t)}}{2A(t)}(Z(t)+f)|\psi_{R}\rangle\right], (44)

which are now time-dependent. In the case of thermal equilibrium, the eigenstates |y1/2|y_{1/2}\rangle reduce to the ground and excited states |ψ0/1|\psi_{0/1}\rangle of the double-well potential.

We recall that in Appendix C it has been shown that the effective density matrix ρ^e\hat{\rho}_{\mathrm{e}} is connected with the reduced density matrix of an average boson ρ^1\hat{\rho}_{1} in a bath of all other N1N-1 bosons via the relation ρ^e=Nρ^1\hat{\rho}_{\mathrm{e}}=N\hat{\rho}_{1}. Thus, the effective density matrix has to obey the physical requirements for the density matrices, in particular, it has to be positive semidefinite, i.e., all its eigenvalues have to be positive. This implies |f|1|f|\leq 1. The ff parameter, therefore, can be interpreted as the pure state fraction in the many-particle bosonic system.

Notice that the term ‘pure state’, so far used to describe an average boson in a pure single-particle state, implies the many-particle Fock state [3]

|Ψ=1N!2N(a^+eiΦb^)N|vac|\Psi\rangle=\frac{1}{\sqrt{N!2^{N}}}(\hat{a}^{\dagger}+e^{i\Phi}\hat{b}^{\dagger})^{N}|vac\rangle (45)

with a^\hat{a}^{\dagger} and b^\hat{b}^{\dagger} being the creation operators relative to the single-particle states ψ0/1\psi_{0/1}, acting on vacuum |vac|vac\rangle. In our framework this state is characterized by a degree of fragmentation 1f=01-f=0. However, it was shown [39, 40] that the state (45) can allow for a nonzero degree of fragmentation due to quantum fluctuations in strongly interacting BECs. A discussion of this regime requires to determine the degree of fragmentation differently, and is not the subject of our study. In particular, the definition used in [39, 40] should not be confused with the one, presented in the paper.

Appendix E Generalised Josephson solutions

Here, we discuss how to obtain the solutions of the generalised Josephson equations (27)-(29).

First, we present the key steps towards the solution for a symmetric double-well. In this case Eq. (29) can be integrated directly and yields A(t)=eBs/cosθ(t)A(t)=e^{B_{\mathrm{s}}}/\cos\theta(t), where BsB_{\mathrm{s}} is an integration constant. Substituting this result into the other two equations, we obtain:

2eBsΔENtanθ(t)\displaystyle\frac{2e^{B_{\mathrm{s}}}\Delta E}{N}\tan\theta(t) =\displaystyle= Z˙(t)\displaystyle-\hbar\dot{Z}(t)
eBsΔEN2Z(t)cos2θ(t)\displaystyle\frac{e^{-B_{\mathrm{s}}}\Delta EN}{2}Z(t)\cos^{2}\theta(t) =\displaystyle= θ˙(t).\displaystyle\hbar\dot{\theta}(t)\,.

Taking the time derivative of the first equation and substituting θ˙(t)\dot{\theta}(t) from the second equation, one obtains an oscillatory equation for the population imbalance Z(t)Z(t):

Z¨(t)+(ΔE)2Z(t)=0.\ddot{Z}(t)+\left(\frac{\Delta E}{\hbar}\right)^{2}Z(t)=0.

Its general solution is Z(t)=Δρssin[ΔEt/+ϕ0s]Z(t)=\Delta\rho_{\mathrm{s}}\sin\left[\Delta Et/\hbar+\phi_{0\mathrm{s}}\right] with the integration constants Δρs\Delta\rho_{\mathrm{s}} and ϕ0s\phi_{0\mathrm{s}}. This gives θ(t)=arctan[NeBsΔρs2cos(ΔEt/+ϕ0s)]\theta(t)=-\arctan\left[\frac{Ne^{-B_{\mathrm{s}}}\Delta\rho_{\mathrm{s}}}{2}\cos\left(\Delta Et/\hbar+\phi_{0\mathrm{s}}\right)\right], which with A(t)=eBs/cosθ(t)A(t)=e^{B_{\mathrm{s}}}/\cos\theta(t) defines the most general solution of the system of the differential equations (27)-(29) with three integration constants BsB_{\mathrm{s}}, Δρs\Delta\rho_{\mathrm{s}} and ϕ0s\phi_{0\mathrm{s}}.

Second, we discuss the case of an asymmetric double-well potential. The integration of Eq. (29) in this case gives A(t)=1cosθ(t)eB+EREL0t𝑑ttanθ(t)A(t)=\frac{1}{\cos\theta(t)}e^{B+\frac{E_{R}-E_{L}}{\hbar}\int_{0}^{t}dt^{\prime}\tan\theta(t^{\prime})}, where BB is an integration constant. Substituting this result into the other two equations, we obtain:

Z˙(t)=4KNtanθ(t)eB+EREL0t𝑑ttanθ(t),\hbar\dot{Z}(t)=\frac{4K}{N}\tan\theta(t)e^{B+\frac{E_{R}-E_{L}}{\hbar}\int_{0}^{t}dt^{\prime}\tan\theta(t^{\prime})},
KNZ(t)eBEREL0t𝑑ttanθ(t)cos2θ(t)=θ˙(t)(EREL).KNZ(t)e^{-B-\frac{E_{R}-E_{L}}{\hbar}\int_{0}^{t}dt^{\prime}\tan\theta(t^{\prime})}\cos^{2}\theta(t)=\\ -\hbar\dot{\theta}(t)-(E_{R}-E_{L})\,.

Taking the time derivative of the first equation and substituting θ˙(t)\dot{\theta}(t) from the second equation, one obtains an oscillatory equation for the population imbalance Z(t)Z(t)

Z¨(t)+(2K)2Z(t)=4K(EREL)N2eB+EREL0t𝑑ttanθ(t).\ddot{Z}(t)+\left(\frac{2K}{\hbar}\right)^{2}Z(t)\\ =-\frac{4K(E_{R}-E_{L})}{N\hbar^{2}}e^{B+\frac{E_{R}-E_{L}}{\hbar}\int_{0}^{t}dt^{\prime}\tan\theta(t^{\prime})}\,.

We see that the Josephson oscillations are governed by two frequencies 2K/=ΔEsin2ξ/2K/\hbar=-\Delta E\sin 2\xi/\hbar and (EREL)/=ΔEcos2ξ/(E_{R}-E_{L})/\hbar=\Delta E\cos 2\xi/\hbar (see Sec. II.2). The θ(t)\theta(t) function can be found by taking time derivative of the second equation and equating the Z˙(t)\dot{Z}(t) from the two equations. This way we obtain

θ¨+(θ˙+EREL)(2θ˙+EREL)tanθ+2K22sin2θ=0.\ddot{\theta}+\left(\dot{\theta}+\frac{E_{R}-E_{L}}{\hbar}\right)\left(2\dot{\theta}+\frac{E_{R}-E_{L}}{\hbar}\right)\tan\theta\\ +\frac{2K^{2}}{\hbar^{2}}\sin 2\theta=0\,.

One can simplify this equation by change of variables y(t)=tanθ(t)y(t)=\tan\theta(t). This gives a nonlinear equation:

y¨+(2K)2y+3ERELyy˙+(EREL)2(1+y2)y=0.\ddot{y}+\left(\frac{2K}{\hbar}\right)^{2}y+3\frac{E_{R}-E_{L}}{\hbar}y\dot{y}\\ +\left(\frac{E_{R}-E_{L}}{\hbar}\right)^{2}(1+y^{2})y=0.

Let us look at the regime of small ELERE_{L}-E_{R}, so the double-well is close to a symmetric one. Then we have 2KΔE2K\approx-\Delta E, EREL=V0E_{R}-E_{L}=-V_{0} (for the latter equality see Sec. II.2) and yysym+uy\approx y_{\mathrm{sym}}+u, where ysymy_{\mathrm{sym}} is a solution in symmetric case. To linear order in uu and V0/ΔEV_{0}/\Delta E we obtain

y¨sym+u¨+(ΔE)2ysym+(ΔE)2u3V0ysymy˙sym=0.\ddot{y}_{\mathrm{sym}}+\ddot{u}+\left(\frac{\Delta E}{\hbar}\right)^{2}y_{\mathrm{sym}}+\left(\frac{\Delta E}{\hbar}\right)^{2}u\\ -\frac{3V_{0}}{\hbar}y_{sym}\dot{y}_{\mathrm{sym}}=0.

In the zeroth order of V0/ΔEV_{0}/\Delta E and small deviation uu one gets the same oscillatory solution ysym(t)y_{\mathrm{sym}}(t) as in the symmetric case, discussed above. In the first order of V0/ΔEV_{0}/\Delta E and small deviation uu we obtain

u¨+(ΔE)2u=3V02ΔE(NeBsΔρs2ΔE)2sin[2ΔEt+2ϕ0s].\ddot{u}+\left(\frac{\Delta E}{\hbar}\right)^{2}u\\ =\frac{3V_{0}}{2\Delta E}\left(\frac{Ne^{-B_{\mathrm{s}}}\Delta\rho_{\mathrm{s}}}{2}\frac{\Delta E}{\hbar}\right)^{2}\sin\left[-2\frac{\Delta Et}{\hbar}+2\phi_{0\mathrm{s}}\right].

The solution of the last equation is

u(t)=(NeBsΔρs2)2V02ΔEsin[2ΔEt+2ϕ0s]+c1V02ΔEsin[ΔEt+ϕ0a]u(t)=-\left(\frac{Ne^{-B_{\mathrm{s}}}\Delta\rho_{\mathrm{s}}}{2}\right)^{2}\frac{V_{0}}{2\Delta E}\sin\left[-\frac{2\Delta Et}{\hbar}+2\phi_{0\mathrm{s}}\right]\\ +c_{1}\frac{V_{0}}{2\Delta E}\sin\left[-\frac{\Delta Et}{\hbar}+\phi_{0\mathrm{a}}\right]

and for θ(t)\theta(t) one gets

θ(t)\displaystyle\theta(t) =\displaystyle= arctan(NeBsΔρs2cosψt)\displaystyle\arctan\left(\frac{Ne^{-B_{\mathrm{s}}}\Delta\rho_{\mathrm{s}}}{2}\cos\psi_{\mathrm{t}}\right)
+1ζ2(t)V02ΔEc1sin[ψt+δϕ0]\displaystyle+\frac{1}{\zeta^{2}(t)}\frac{V_{0}}{2\Delta E}\,c_{1}\sin\left[\psi_{\mathrm{t}}+\delta\phi_{0}\right]
1ζ2(t)V02ΔE(NeBsΔρs2)2sin[2ψt].\displaystyle-\frac{1}{\zeta^{2}(t)}\frac{V_{0}}{2\Delta E}\left(\frac{Ne^{-B_{\mathrm{s}}}\Delta\rho_{\mathrm{s}}}{2}\right)^{2}\sin\left[2\psi_{\mathrm{t}}\right].

This solution depends on the integration constants BsB_{s}, Δρs\Delta\rho_{s} and ϕ0s\phi_{0s} of the symmetric double-well problem (our zeroth order in V0/ΔEV_{0}/\Delta E solution) and on the two new integration constants ϕ0a\phi_{0a} and c1c_{1}. Taking into account that the other integration constant BB should only slightly differ from the analogous constant of the symmetric double-well problem B=Bs+V0/(2ΔE)δBB=B_{\mathrm{s}}+V_{0}/(2\Delta E)\delta B we obtain the solution for A(t)A(t)

A(t)=eBsζ(t)+V02ΔENΔρsζ(t)(sinψtsinϕ0s)V02ΔEe2Bsζ(t)(ΔρsN2)3cosψtsin(2ψt)+V02ΔEc1ΔρsN2ζ(t)cosψtsin(δϕ0+ψt)+V02ΔEeBsζ(t)δB.A(t)=e^{B_{\mathrm{s}}}\zeta(t)+\frac{V_{0}}{2\Delta E}N\Delta\rho_{\mathrm{s}}\zeta(t)\left(\sin\psi_{\mathrm{t}}-\sin\phi_{0\mathrm{s}}\right)\\ -\frac{V_{0}}{2\Delta E}\frac{e^{-2B_{\mathrm{s}}}}{\zeta(t)}\left(\frac{\Delta\rho_{\mathrm{s}}N}{2}\right)^{3}\cos\psi_{\mathrm{t}}\sin\left(2\psi_{\mathrm{t}}\right)\\ +\frac{V_{0}}{2\Delta E}c_{1}\frac{\Delta\rho_{\mathrm{s}}N}{2\zeta(t)}\cos\psi_{\mathrm{t}}\sin\left(\delta\phi_{0}+\psi_{\mathrm{t}}\right)+\frac{V_{0}}{2\Delta E}e^{B_{\mathrm{s}}}\zeta(t)\delta B. (47)

The constants ϕ0a\phi_{0\mathrm{a}}, δB\delta B and c1c_{1} can be fixed by the initial condition for the effective density matrix, which for the slightly asymmetric double-well is as well a linear function of a small parameter V0/ΔEV_{0}/\Delta E. Thus, for fractional imbalance we have

Z(t)=[ΔρsV02ΔEeBsN((ΔρsN)2sinϕ0s2c1e2Bssinδϕ0eBsδBΔρsN)]sinψt2V0ΔEeBsNV0ΔEeBsNc1cosδϕ0cosψt,Z(t)=\left[\Delta\rho_{\mathrm{s}}-\frac{V_{0}}{2\Delta E}\frac{e^{-B_{\mathrm{s}}}}{N}\left((\Delta\rho_{\mathrm{s}}N)^{2}\sin\phi_{0\mathrm{s}}\right.\right.\\ \left.\left.-2c_{1}e^{2B_{\mathrm{s}}}\sin\delta\phi_{0}-e^{B_{\mathrm{s}}}\delta B\Delta\rho_{\mathrm{s}}N\right)\right]\sin\psi_{\mathrm{t}}\\ -2\frac{V_{0}}{\Delta E}\frac{e^{B_{\mathrm{s}}}}{N}-\frac{V_{0}}{\Delta E}\frac{e^{B_{\mathrm{s}}}}{N}c_{1}\cos\delta\phi_{0}\cos\psi_{\mathrm{t}}, (48)

where the denotation 1+(NeBsΔρs2)2cos2ψt=ζ(t)\sqrt{1+\left(\frac{Ne^{-B_{\mathrm{s}}}\Delta\rho_{\mathrm{s}}}{2}\right)^{2}\cos^{2}\psi_{\mathrm{t}}}=\zeta(t) was used.

The generalized solution reduces to the standard one, when the pure state condition A(t)=N1Z(t)2/2A(t)=N\sqrt{1-Z(t)^{2}}/2 is satisfied. It holds for all times tt when the integration constants in the case of the symmetric double-well obey the relation 1Δρs2=(2/N)2e2Bs1-\Delta\rho_{\mathrm{s}}^{2}=(2/N)^{2}e^{2B_{\mathrm{s}}}. For the asymmetric double-well, in the first order of V0/ΔEV_{0}/\Delta E, the pure state condition for the parameters of the generalized solution reads

2c1e2BsΔρssinδϕ+Nf(s)(δBeBsNΔρssinϕ0s)=02c_{1}e^{2B_{\mathrm{s}}}\Delta\rho_{\mathrm{s}}\sin\delta\phi+Nf^{\mathrm{(s)}}\left(\delta Be^{B_{\mathrm{s}}}-N\Delta\rho_{\mathrm{s}}\sin\phi_{0\mathrm{s}}\right)=0

In the first order of V0/ΔEV_{0}/\Delta E the generalized and the pure state solutions are perfectly consistent for the following choice of constants of generalized solution:

δB\displaystyle\delta B =\displaystyle= 2(Δρssinϕ0sc2)1Δρs2,\displaystyle\frac{2(\Delta\rho_{\mathrm{s}}\sin\phi_{0\mathrm{s}}-c_{2})}{\sqrt{1-\Delta\rho_{\mathrm{s}}^{2}}},
c1\displaystyle c_{1} =\displaystyle= 2αΔρs1Δρs21Δρs2+c22α2Δρs4,\displaystyle\frac{2\alpha\Delta\rho_{\mathrm{s}}}{1-\Delta\rho_{\mathrm{s}}^{2}}\sqrt{1-\Delta\rho_{\mathrm{s}}^{2}+\frac{c_{2}^{2}}{\alpha^{2}\Delta\rho_{\mathrm{s}}^{4}}},
δϕ0\displaystyle\delta\phi_{0} =\displaystyle= arctan[c2αΔρs21Δρs2].\displaystyle\arctan\left[\frac{c_{2}}{\alpha\Delta\rho_{\mathrm{s}}^{2}\sqrt{1-\Delta\rho_{\mathrm{s}}^{2}}}\right].

References

  • Pethick and Smith [2008] C. J. Pethick and H. Smith, Bose–Einstein condensation in dilute gases (Cambridge university press, 2008).
  • Ketterle [2002] W. Ketterle, Reviews of Modern Physics 74, 1131 (2002).
  • Pitaevskii and Stringari [2016] L. Pitaevskii and S. Stringari, Bose-Einstein condensation and superfluidity, Vol. 164 (Oxford University Press, 2016).
  • Anderson et al. [1995] M. H. Anderson, J. R. Ensher, M. R. Matthews, C. E. Wieman, and E. A. Cornell, science 269, 198 (1995).
  • Davis et al. [1995] K. B. Davis, M.-O. Mewes, M. R. Andrews, N. J. van Druten, D. S. Durfee, D. M. Kurn, and W. Ketterle, Physical review letters 75, 3969 (1995).
  • Unnikrishnan [2018] C. Unnikrishnan, Bose-Einstein condensates as universal quantum matter (2018).
  • Smerzi et al. [1997] A. Smerzi, S. Fantoni, S. Giovanazzi, and S. R. Shenoy, Phys. Rev. Lett. 79, 4950 (1997).
  • Bloch et al. [2008] I. Bloch, J. Dalibard, and W. Zwerger, Reviews of modern physics 80, 885 (2008).
  • Bloch [2005] I. Bloch, Nature physics 1, 23 (2005).
  • Gross [1961] E. P. Gross, Il Nuovo Cimento (1955-1965) 20, 454 (1961).
  • Pitaevskii [1961] L. P. Pitaevskii, Sov. Phys. JETP 13, 451 (1961).
  • Leggett [1999] A. J. Leggett, Rev. Mod. Phys. 71, S318 (1999).
  • Ghosh et al. [2019] S. Ghosh, J. Bera, P. K. Panigrahi, and U. Roy, International Journal of Quantum Information 17, 1950019 (2019).
  • Chiow and Yu [2023] S.-w. Chiow and N. Yu, in Quantum Sensing, Imaging, and Precision Metrology (SPIE, 2023) p. PC124471W.
  • Ockeloen et al. [2013] C. F. Ockeloen, R. Schmied, M. F. Riedel, and P. Treutlein, Physical review letters 111, 143001 (2013).
  • Edwards [2013] M. Edwards, Nature Physics 9, 68 (2013).
  • Sewell et al. [2010] R. Sewell, J. Dingjan, F. Baumgärtner, I. Llorente-García, S. Eriksson, E. A. Hinds, G. Lewis, P. Srinivasan, Z. Moktadir, C. O. Gollasch, et al., Journal of Physics B: Atomic, Molecular and Optical Physics 43, 051003 (2010).
  • Rudolph et al. [2015] J. Rudolph, W. Herr, C. Grzeschik, T. Sternke, A. Grote, M. Popp, D. Becker, H. Müntinga, H. Ahlers, A. Peters, et al., New Journal of Physics 17, 065001 (2015).
  • Leggett [2001] A. J. Leggett, Rev. Mod. Phys. 73, 307 (2001).
  • Dalfovo et al. [1999] F. Dalfovo, S. Giorgini, L. P. Pitaevskii, and S. Stringari, Rev. Mod. Phys. 71, 463 (1999).
  • Salasnich et al. [1999] L. Salasnich, A. Parola, and L. Reatto, Physical Review A 60, 4171 (1999).
  • Cordes and Das [2001] J. Cordes and A. Das, Superlattices and microstructures 29, 121 (2001).
  • Milburn et al. [1997] G. J. Milburn, J. Corney, E. M. Wright, and D. F. Walls, Phys. Rev. A 55, 4318 (1997).
  • Hasegawa [2013] H. Hasegawa, Physica A: Statistical Mechanics and its Applications 392, 6232 (2013).
  • Song [2008] D.-Y. Song, Annals of Physics 323, 2991 (2008).
  • Wang et al. [2007] W. Wang, L. B. Fu, and X. X. Yi, Physical Review A 75, 045601 (2007).
  • Bidasyuk et al. [2018] Y. M. Bidasyuk, M. Weyrauch, M. Momme, and O. O. Prikhodko, Journal of Physics B: Atomic, Molecular and Optical Physics 51, 205301 (2018).
  • Cataldo and Jezek [2014] H. M. Cataldo and D. M. Jezek, Phys. Rev. A 90, 043610 (2014).
  • Albiez et al. [2005] M. Albiez, R. Gati, J. Fölling, S. Hunsmann, M. Cristiani, and M. K. Oberthaler, Physical Review Letters 95, 010402 (2005).
  • Fattori et al. [2008] M. Fattori, C. D’Errico, G. Roati, M. Zaccanti, M. Jona-Lasinio, M. Modugno, M. Inguscio, and G. Modugno, Physical review letters 100, 080405 (2008).
  • Binney and Skinner [2013] J. Binney and D. Skinner, The physics of quantum mechanics (Oxford University Press, 2013).
  • Dauphinee and Marsiglio [2015] T. Dauphinee and F. Marsiglio, American Journal of Physics 83, 861 (2015).
  • Mazets and Schmiedmayer [2010] I. Mazets and J. Schmiedmayer, New Journal of Physics 12, 055023 (2010).
  • Landau and Lifshitz [2013] L. D. Landau and E. M. Lifshitz, Statistical Physics: Volume 5, Vol. 5 (Elsevier, 2013).
  • Gustafson and Sigal [2003] S. J. Gustafson and I. M. Sigal, Density matrices, in Mathematical Concepts of Quantum Mechanics (Springer Berlin Heidelberg, Berlin, Heidelberg, 2003).
  • Erdahl and Smith Jr [2012] R. M. Erdahl and V. H. Smith Jr, Density matrices and density functionals: proceedings of the A. John Coleman symposium (Springer Science & Business Media, 2012).
  • Mueller et al. [2006] E. J. Mueller, T.-L. Ho, M. Ueda, and G. Baym, Phys. Rev. A 74, 033612 (2006).
  • Spekkens and Sipe [1999] R. W. Spekkens and J. E. Sipe, Phys. Rev. A 59, 3868 (1999).
  • Pitaevskii and Stringari [2001] L. Pitaevskii and S. Stringari, Physical Review Letters 87, 180402 (2001).
  • Gati et al. [2006] R. Gati, B. Hemmerling, J. Fölling, M. Albiez, and M. K. Oberthaler, Phys. Rev. Lett. 96, 130404 (2006).
  • Salasnich and Tiene [2016] L. Salasnich and A. Tiene, Dissipative effects in the dynamics of a Bose-Einstein condensate under double-well confinement (2016).
  • Spagnolli et al. [2017] G. Spagnolli, G. Semeghini, L. Masi, G. Ferioli, A. Trenkwalder, S. Coop, M. Landini, L. Pezzè, G. Modugno, M. Inguscio, A. Smerzi, and M. Fattori, Phys. Rev. Lett. 118, 230403 (2017).
  • Cronin et al. [2009] A. D. Cronin, J. Schmiedmayer, and D. E. Pritchard, Reviews of Modern Physics 81, 1051 (2009).
  • Smalley et al. [1977] R. E. Smalley, L. Wharton, and D. H. Levy, Accounts of Chemical Research 10, 139 (1977).
  • Phillips et al. [1991] W. D. Phillips, P. D. Lett, S. Rolston, C. Tanner, R. Watts, C. Westbrook, C. Salomon, J. Dalibard, A. Clairon, and S. Guellati, Physica Scripta 1991, 20 (1991).
  • Prentiss et al. [1989] M. Prentiss, A. Cable, and N. Bigelow, JOSA B 6, 2155 (1989).