Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
License: arXiv.org perpetual non-exclusive license
arXiv:2311.00540v2 [cond-mat.supr-con] 19 Mar 2024

Critical current throughout the BCS-BEC crossover
with the inclusion of pairing fluctuations

L. Pisani School of Science and Technology, Physics Division, Università di Camerino, 62032 Camerino (MC), Italy    V. Piselli CNR-INO, Istituto Nazionale di Ottica, Sede di Firenze, 50125 (FI), Italy    G. Calvanese Strinati giancarlo.strinati@unicam.it School of Science and Technology, Physics Division, Università di Camerino, 62032 Camerino (MC), Italy CNR-INO, Istituto Nazionale di Ottica, Sede di Firenze, 50125 (FI), Italy
Abstract

The present work aims at providing a systematic analysis of the current density versus momentum characteristics for a fermionic superfluid throughout the BCS-BEC crossover, even in the fully homogeneous case. At low temperatures, where pairing fluctuations are not strong enough to invalidate a quasi-particle approach, a sharp threshold for the inception of a back-flow current is found, which sets the onset of dissipation and identifies the critical momentum according to Landau. This momentum is seen to smoothly evolve from the BCS to the BEC regimes, whereby a single expression for the single-particle current density that includes pairing fluctuations enables us to incorporate on equal footing two quite distinct dissipative mechanisms, namely, pair-breaking and phonon excitations in the two sides of the BCS-BEC crossover, respectively. At finite temperature, where thermal fluctuations broaden the excitation spectrum and make the dissipative (kinetic and thermal) mechanisms intertwined with each other, an alternative criterion due to Bardeen is instead employed to signal the loss of superfluid behavior. In this way, detailed comparison with available experimental data in linear and annular geometries is significantly improved with respect to previous approaches, thereby demonstrating the crucial role played by quantum fluctuations in renormalizing the single-particle excitation spectrum.

I Introduction

The concepts of the critical current of a superconductor Tinkham-1996 and the critical velocity of a superfluid LL-1980 are intimately related. In both cases, the breaking of dissipation-less flow occurs when the imparted kinetic energy gives rise to quasi-particle excitations with zero energy. For superconductors, this shows up in the phenomenon known as “gapless superconductivity” Fulde-1965 ; Parmenter-1965 after the seminal work by Abrikosov and Gor’kov on the effect of impurities in a superconductor Abrikosov-1960 , which was later generalized by Maki Parks-1969 to the case of a time-reversal symmetry breaking (de-pairing or pair-breaking) agent Fulde-1963 . For superfluids, it corresponds to the criterion for the onset of viscous flow originally conceived by Landau for 44{}^{4}start_FLOATSUPERSCRIPT 4 end_FLOATSUPERSCRIPTHe LL-1980 . To the extent that the properties of superconductors and superfluids crucially depend on the details of the underlying energy spectrum, a detailed analysis of the onset of the dissipative flow should be able to reveal the microscopic mechanisms underlying this phenomenon.

The appropriate physical quantity for revealing the above features is the current response to an overall momentum imposed on the (superconducting/superfluid) system. The behavior of this physical quantity is familiar in two paradigmatic cases, namely, the weakly-attractive Fermi gas and the weakly-repulsive Bose gas. The first one can be described in terms of the Bardeen-Cooper-Schrieffer (BCS) theory of conventional superconductors Bardeen-1962 (with its extension to dirty superconductors Kupri-1980 ), while the second one in terms of the Bogoliubov theory of weakly-interacting bosons Fetter-1972 that Bose-Einstein condense (BEC) at low temperature (with its extension to strongly-interacting bosons like 44{}^{4}start_FLOATSUPERSCRIPT 4 end_FLOATSUPERSCRIPTHe Kramer-1969 ). In both systems, kinetic and/or thermal dissipation occurs in the form of a back-flow current (which is somewhat analogous to that occurring in a normal Fermi system Nozieres-1997 ), giving rise to density inhomogeneities which, in turn, lead to the formation of vortices Tinkham-1996 as well as of rotons Feynman-1956 .

With the advent of the ultra-cold atomic gases (and, specifically, of Fermi gases with attractive inter-particle interaction), these two paradigmatic (fermionic and bosonic) physical systems can be smoothly connected via Fano-Feshbach resonances, whereby these systems correspond to the limiting regimes of the so-called BCS-BEC crossover, with largely-overlapping Cooper pairs in the BCS regime smoothly evolving into tightly-bound composite bosons in the BEC regime. (A recent comprehensive review on the BCS-BEC crossover can be found in Ref. Physics-Reports-2018 .) Commonly, the BCS-BEC crossover is spanned in terms of the dimensionless coupling (kFaF)1superscriptsubscript𝑘𝐹subscript𝑎𝐹1(k_{F}a_{F})^{-1}( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, where kF=(3π2n)1/3subscript𝑘𝐹superscript3superscript𝜋2𝑛13k_{F}=(3\pi^{2}n)^{1/3}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = ( 3 italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_n ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT is the Fermi momentum with density n𝑛nitalic_n and aFsubscript𝑎𝐹a_{F}italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT the scattering length of the two-fermion problem in vacuum. This coupling ranges from (kFaF)11less-than-or-similar-tosuperscriptsubscript𝑘𝐹subscript𝑎𝐹11(k_{F}\,a_{F})^{-1}\lesssim-1( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ≲ - 1 in the weak-coupling (BCS) regime when aF<0subscript𝑎𝐹0a_{F}<0italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT < 0, to (kFaF)1+1greater-than-or-equivalent-tosuperscriptsubscript𝑘𝐹subscript𝑎𝐹11(k_{F}\,a_{F})^{-1}\gtrsim+1( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ≳ + 1 in the strong-coupling (BEC) regime when aF>0subscript𝑎𝐹0a_{F}>0italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT > 0, passing through the unitary limit (kFaF)1=0superscriptsubscript𝑘𝐹subscript𝑎𝐹10(k_{F}\,a_{F})^{-1}=0( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = 0 when |aF|subscript𝑎𝐹|a_{F}|| italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT | diverges.

Experiments with ultra-cold Fermi gases, with their detailed control on the relevant parameters associated with the degrees of freedom of the system Hamiltonian, have revived the interest in many key aspects of superconductivity (or, more generally, of fermionic superfluidity), including the Abrikosov vortex lattice Zwierlein-2005 ; Simonucci-2015 , the Josephson effect Miller-2007 ; Spuntarelli-2007 , and the interaction of magnetic(-like) fields with fermion spins Zwierlein-2006 ; Partridge-2006 ; Pieri-2006 . At the same time, these experiments have also stimulated more refined and advanced theoretical approaches addressing these aspects, inevitably promoting a valuable feedback in the theory of superconductors themselves. The topic dealt with in the present article, about the current response to an overall momentum imposed on a (homogeneous) fermionic superfluid, makes no exception to this close connection between experiment and theory which is feasible with ultra-cold Fermi gases. In particular, two recent experiments have measured at low temperature throughout the BCS-BEC crossover, both the critical velocity in a linear geometry where a weak barrier procedes through the superfluid in a circular motion Weimer-2015 , and the maximum quantized persistent current circulating in an annular geometry DelPace-2022 . Both experiments were essentially aimed at identifying the value of the critical current (or velocity) for an ultra-cold Fermi gas, in a way that would resemble as closely as possible the Landau criterion envisioned theoretically for liquid helium Landau-1941 .

In this context, the present article considers the smooth evolution of the current-vs-momentum characteristics from the BCS to the BEC limits of the BCS-BEC crossover at any temperature in the superfluid phase, resting on an approach recently developed in Ref. Pisani-2023 to deal with the effects of pairing fluctuations in the presence of a supercurrent and nontrivial spatial constraints on equal footing (although in the present case the spatial constraint may at most correspond to a weak barrier in line with the setups of the experiments of Refs. Weimer-2015 ; DelPace-2022 mentioned above). This theoretical approach will enable us to demonstrate, in terms of a single theory, how the Landau critical velocity at zero temperature, originally introduced for a bosonic superfluid Landau-1941 , smoothly evolves into its counterpart in the BCS regime where, however, the presence of an underlying Fermi surface makes the value of the critical velocity different from that obtained by the Bardeen criterion originally introduced for a fermionic superfluid Bardeen-1962 . In addition, the analysis of the current-vs-momentum characteristics obtained in this way will enable us to identify the critical current (or, else, the critical velocity) even at finite temperature in terms of the Bardeen criterion.

The results for the current-momentum response obtained by the present approach, which include the renormalization of the single-particle excitation spectrum due to pairing fluctuations, represents a definite improvement over those given by the conventional BCS-RPA (Random Phase Approximation) theory of superconductors, originally introduced by Anderson for conventional superconductors Anderson-1958 and later extended to the BCS-BEC crossover MPS-1998 ; Comb-2006 ; Spunta-2010 .

The main results obtained in this article are as follows:

(i) The current-vs-momentum characteristics are obtained with the inclusion of pairing fluctuations beyond mean field for a homogeneous two-component Fermi gas with attractive inter-particle interaction, at any temperature in the superfluid phase and throughout the BCS-BEC crossover. Although, in principle, the value of the critical current obtained in this way is equivalent to that due to an applied infinitesimal perturbation, in practice the presence of a non negligible impurity (in the form of a small barrier) embedded in an otherwise homogeneous superfluid has to be taken into account when simulating realistic experimental setups. This will be done in terms of the theoretical mLPDA approach (with the acronym standing for modified Local Phase Density Approximation) that was recently developed in Ref. Pisani-2023 , with the further consideration of the extended Gorkov-Melik-Barkhudarov (GMB) approach implemented in Ref. Pisani-2018-b that improves on the comparison with experimental data.

(ii) A rather good comparison is retrieved in this way with the experimental data available at low temperature, specifically, from Ref. Weimer-2015 for the critical velocity in a linear geometry and from Ref. DelPace-2022 for the maximum value of the quantized velocity in an annular geometry.

(iii) The concept of the intrinsic critical current is further extended locally inside a realistic barrier, and shown to account for the value of the critical current which is involved in the Josephson effect at finite temperature as obtained in Ref. Pisani-2023 .

(iv) Overall, this analysis enables us to assess how the Landau and Bardeen criteria manifest themselves in different experimental contexts and at the relevant temperature, when the coupling is varied across the BCS-BEC crossover.

The article is organized as follows. In Section II the current-vs-momentum characteristics are considered for the simple cases of a weakly-attractive Fermi gas and of a weakly-repulsive Bose gas, that represent the limiting cases occurring in the BCS-BEC crossover. In Section III the inclusion of pairing fluctuations beyond mean field will allow us to obtain a continuous evolution of the current-vs-momentum characteristics for a homogeneous Fermi system that evolves from the BCS to the BEC regimes, thus making the intrinsic critical current obtained in this way to be the theoretical benchmark for interpreting microscopically experiments and theoretical simulations. A linear geometry is first considered to relate with the experimental results on the critical velocity of Ref. Weimer-2015 across the BCS-BEC crossover as well as with the theoretical results on the Josephson characteristics obtained in Ref. Pisani-2023 . Section IV considers alternatively an annular geometry with quantized values of the superfluid velocity, for which comparison with the experimental data of Ref. DelPace-2022 is possible. Section V gives our conclusions.

Finally, for the benefit of the readers Appendix A briefly summarizes previous theoretical results which are utilized for the specific purposes of the present work, while Appendix B expands on the topics dealt with in Secs. II.1 and III.2.

II Intrinsic critical current of weakly-interacting gases

This Section briefly reviews the current-density response induced by an imposed momentum in two paradigmatic cases: The weakly-attractive Fermi gas treated within the BCS approximation and the weakly-repulsive Bose gas treated within the Bogoliubov approximation. In the Bogoliubov case, the Landau critical velocity sets not only the dissipative threshold but also the onset of the dynamical instability of the gas. In the BCS case, on the other hand, the dissipative threshold does not coincide with the onset of the dynamical instability. This is due to the presence of an underlying Fermi surface which introduces a second (thermodynamic) critical velocity, that can be referred to as the Bardeen critical velocity after Bardeen who first considered it Bardeen-1962 .

The expressions of the current density utilized here for the Fermi and Bose gases can be obtained as limiting cases of a more general expression which spans the BCS-BEC crossover between the two (BCS and BEC) limits, which will be extensively discussed in Sec. III below. In addition, for the benefit of the readers Appendix A summarizes how this more general expression for the current was originally obtained in Ref. Pisani-2023 .

II.1 Intrinsic critical current of a
       weakly-attractive Fermi gas

We first consider the superfluid current density of a weakly-attractive Fermi gas as a function of an imposed velocity, which we calculate at the mean-field (BCS) level. Although this topic has already been dealt with by a number of authors Bardeen-1962 ; Parks-1969 ; Fulde-1963 ; Hansen-1968 in the context of de-pairing currents and gapless superconductivity we are here interested in its intimate relation with the system energy excitation spectrum and with the Landau criterion for superfluidity Khalatnikov-2000 .

At the mean-field level, the current density induced by an imposed momentum 𝐪𝐪{\bf q}bold_q reads Piselli-2020

𝐣(𝐪)=n𝐪m+2d𝐤(2π)3𝐤mf(E+(𝐤;𝐪)),𝐣𝐪𝑛𝐪𝑚2𝑑𝐤superscript2𝜋3𝐤𝑚𝑓subscript𝐸𝐤𝐪{\mathbf{j}}({\bf q})=n{{\bf q}\over m}+2\!\int\!{d{\bf k}\over(2\pi)^{3}}\,{{% \bf k}\over m}\;f(E_{+}({\bf k};{\bf q}))\,,bold_j ( bold_q ) = italic_n divide start_ARG bold_q end_ARG start_ARG italic_m end_ARG + 2 ∫ divide start_ARG italic_d bold_k end_ARG start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG divide start_ARG bold_k end_ARG start_ARG italic_m end_ARG italic_f ( italic_E start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ( bold_k ; bold_q ) ) , (1)

where n𝑛nitalic_n is the (fermionic) particle density, m𝑚mitalic_m the fermion mass, E+(𝐤;𝐪)=E(𝐤;𝐪)+𝐤𝐪msubscript𝐸𝐤𝐪𝐸𝐤𝐪𝐤𝐪𝑚E_{+}({\bf k};{\bf q})=E({\bf k};{\bf q})+{{\bf k}\cdot{\bf q}\over m}italic_E start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ( bold_k ; bold_q ) = italic_E ( bold_k ; bold_q ) + divide start_ARG bold_k ⋅ bold_q end_ARG start_ARG italic_m end_ARG the Doppler-shifted BCS quasi-particle spectrum where E(𝐤;𝐪)=(𝐤22mμ+𝐪22m)2+Δ𝐪2𝐸𝐤𝐪superscriptsuperscript𝐤22𝑚𝜇superscript𝐪22𝑚2superscriptsubscriptΔ𝐪2E({\bf k};{\bf q})=\sqrt{({\mathbf{k}^{2}\over 2m}-\mu+{{\bf q}^{2}\over 2m})^% {2}+\Delta_{{\bf q}}^{2}}italic_E ( bold_k ; bold_q ) = square-root start_ARG ( divide start_ARG bold_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG - italic_μ + divide start_ARG bold_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Δ start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG with gap parameter Δ𝐪subscriptΔ𝐪\Delta_{{\bf q}}roman_Δ start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT and chemical potential μ𝜇\muitalic_μ, f(ϵ)=(eϵkBT+1)1𝑓italic-ϵsuperscriptsuperscript𝑒italic-ϵsubscript𝑘𝐵𝑇11f(\epsilon)=(e^{\epsilon\over k_{B}T}+1)^{-1}italic_f ( italic_ϵ ) = ( italic_e start_POSTSUPERSCRIPT divide start_ARG italic_ϵ end_ARG start_ARG italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG end_POSTSUPERSCRIPT + 1 ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT the Fermi distribution function, and 2222 the spin factor. [We set =1Planck-constant-over-2-pi1\hbar=1roman_ℏ = 1 throughout.]

Here, as well as in the more general expressions (7) and (11), the first term n𝐪m𝑛𝐪𝑚n{{\bf q}\over m}italic_n divide start_ARG bold_q end_ARG start_ARG italic_m end_ARG represents the total density n𝑛nitalic_n moving uniformly in stationary equilibrium with an imposed momentum q=mv𝑞𝑚𝑣q=mvitalic_q = italic_m italic_v, while the second term corresponds to the depletion of the superfluid component in favor of the normal one due either to thermal or velocity effects. This connects with the Landau two-fluid model, as discussed below for both fermions and bosons.

By symmetry consideration, the second term on the right-hand side of Eq. (1) is directed along -𝐪𝐪{\bf q}bold_q and has accordingly the microscopic interpretation of a back-flow current Nozieres-1997 , which is set in by the produced excitations that eventually make the system to dissipate. The mechanism of the back-flow is actually a widely ranging concept, which is invoked in the stability of vortices in superconductors Tinkham-1996 and in the creation of rotons in 44{}^{4}start_FLOATSUPERSCRIPT 4 end_FLOATSUPERSCRIPTHe (as proposed in the seminal work by Feynman Feynman-1956 ). In this respect, the typical return flow of rotons is not a specific feature of a bosonic system like 44{}^{4}start_FLOATSUPERSCRIPT 4 end_FLOATSUPERSCRIPTHe, since it was also detected in a two-dimensional Fermi liquid like 33{}^{3}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPTHe Godfrin-2012 as well as in a dipolar bosonic quantum gas Ferlaino-2018 .

For small momenta qkFmuch-less-than𝑞subscript𝑘𝐹q\ll k_{F}italic_q ≪ italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT, the expression (1) recovers the Landau’s two-fluid model whereby the back-flow term implicitly defines the normal fluid density at finite temperature LL-1980

ρn(T)=2d𝐤(2π)3(𝐤𝐪^)2m(df(E(𝐤))dE(𝐤))subscript𝜌𝑛𝑇2𝑑𝐤superscript2𝜋3superscript𝐤^𝐪2𝑚𝑑𝑓𝐸𝐤𝑑𝐸𝐤\rho_{n}(T)=2\!\int\!{d{\bf k}\over(2\pi)^{3}}\;{({\bf k}\cdot\hat{{\bf q}})^{% 2}\over m}\left(-{df(E({\bf k}))\over dE({\bf k})}\right)italic_ρ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_T ) = 2 ∫ divide start_ARG italic_d bold_k end_ARG start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG divide start_ARG ( bold_k ⋅ over^ start_ARG bold_q end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_m end_ARG ( - divide start_ARG italic_d italic_f ( italic_E ( bold_k ) ) end_ARG start_ARG italic_d italic_E ( bold_k ) end_ARG ) (2)

where E(𝐤)=E(𝐤;𝐪=0)𝐸𝐤𝐸𝐤𝐪0E({\bf k})=E({\bf k};{\bf q}=0)italic_E ( bold_k ) = italic_E ( bold_k ; bold_q = 0 ). In this way, the standard hydrodynamic relation 𝐣(𝐪)=n𝐪mρn(T)𝐪m=ρs(T)𝐪m𝐣𝐪𝑛𝐪𝑚subscript𝜌𝑛𝑇𝐪𝑚subscript𝜌𝑠𝑇𝐪𝑚{\mathbf{j}}({\bf q})=n{{\bf q}\over m}-\rho_{n}(T){{\bf q}\over m}=\rho_{s}(T% ){{\bf q}\over m}bold_j ( bold_q ) = italic_n divide start_ARG bold_q end_ARG start_ARG italic_m end_ARG - italic_ρ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_T ) divide start_ARG bold_q end_ARG start_ARG italic_m end_ARG = italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ) divide start_ARG bold_q end_ARG start_ARG italic_m end_ARG is obtained, where ρs(T)=nρn(T)subscript𝜌𝑠𝑇𝑛subscript𝜌𝑛𝑇\rho_{s}(T)=n-\rho_{n}(T)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ) = italic_n - italic_ρ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_T ) is the superfluid density at temperature T𝑇Titalic_T. In the present work, we go beyond the linear regime of Eq. (2) and focus especially on the non-linear super-critical effects that show up once the Landau dissipation threshold is reached Khalatnikov-2000 .

Refer to caption
Figure 1: Momentum dependence of the current j(q)𝑗𝑞j(q)italic_j ( italic_q ) for a weakly-attractive Fermi gas with coupling (kFaF)1=1.0superscriptsubscript𝑘𝐹subscript𝑎𝐹11.0(k_{F}a_{F})^{-1}=-1.0( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = - 1.0, at several temperatures in the superfluid phase. Here, q𝑞qitalic_q is in units of the Fermi momentum kF=(3π2n)1/3subscript𝑘𝐹superscript3superscript𝜋2𝑛13k_{F}=\left(3\pi^{2}n\right)^{1/3}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = ( 3 italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_n ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT and j(q)𝑗𝑞j(q)italic_j ( italic_q ) in units of the Fermi current JF=nkF/msubscript𝐽𝐹𝑛subscript𝑘𝐹𝑚J_{F}=nk_{F}/mitalic_J start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = italic_n italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT / italic_m. Main panel: At given temperature T𝑇Titalic_T (here in units of the superfluid critical temperature Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT calculated at the mean-filed level), the value JBsubscript𝐽𝐵J_{B}italic_J start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT of the critical current is reached at the maximum of each characteristic in agreement with the Bardeen criterion (see the text). Inset: At T=0𝑇0T=0italic_T = 0, the onset of dissipation is defined by qLsubscript𝑞𝐿q_{L}italic_q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT according to the Landau criterion, while at qBsubscript𝑞𝐵q_{B}italic_q start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT the stability is lost according to the Bardeen criterion.

Figure 1 shows the momentum dependence of the current density given by the expression (1) at several temperatures. In particular, at zero temperature the inset of Fig. 1 shows that the current is proportional to the momentum q𝑞qitalic_q and the back-flow term of Eq. (1) is zero as long as q𝑞qitalic_q satisfies the Landau condition

E(𝐤;𝐪)kqm0q2mqL2mΔ2+μ2μ𝐸𝐤𝐪𝑘𝑞𝑚0superscript𝑞2𝑚superscriptsubscript𝑞𝐿2𝑚superscriptΔ2superscript𝜇2𝜇E({\bf k};{\bf q})-{kq\over m}\geq 0\;\Rightarrow\;{q^{2}\over m}\leq{q_{L}^{2% }\over m}\equiv\sqrt{\Delta^{2}+\mu^{2}}-\mustart_ROW start_CELL italic_E ( bold_k ; bold_q ) - divide start_ARG italic_k italic_q end_ARG start_ARG italic_m end_ARG ≥ 0 ⇒ divide start_ARG italic_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_m end_ARG ≤ divide start_ARG italic_q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_m end_ARG ≡ square-root start_ARG roman_Δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_μ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - italic_μ end_CELL end_ROW (3)

which identifies the limiting (Landau) momentum qLsubscript𝑞𝐿q_{L}italic_q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT. Past qLsubscript𝑞𝐿q_{L}italic_q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT, zero-energy excitations are produced in the system according to the Landau criterion, with the effect of decreasing the strength of the current in a gradual way up to a second critical (Bardeen) momentum qBsubscript𝑞𝐵q_{B}italic_q start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT. At this point, the current reaches its maximum value which signals the limit of the thermodynamic stability of the system. Past qBsubscript𝑞𝐵q_{B}italic_q start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT, in fact, the dissipation becomes out of control in such a way that the thermodynamic equilibrium is lost and a new phase sets in. Accordingly, at zero temperature the onset of dynamical dissipation occurring at the Landau momentum qLsubscript𝑞𝐿q_{L}italic_q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT can be identified when the current-vs-momentum characteristic turns from a linear to a non-linear behavior, while the thermodynamic stability is lost at a second critical momentum qBsubscript𝑞𝐵q_{B}italic_q start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT when the characteristic attains its maximum value. A rigorous definition of qBsubscript𝑞𝐵q_{B}italic_q start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT will be given after the expression (4) below when discussing the case of finite temperature.

At finite temperature, on the other hand, kinetic and thermal dissipations are simultaneously present making the identification of qLsubscript𝑞𝐿q_{L}italic_q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT from the current-vs-momentum characteristic no longer possible, since a sharp transition from a linear to a non-linear behavior is now lost (as shown in the main panel of Fig. 1). In this case, it is convenient to adopt a thermodynamic criterion for the stability of the system against the collapse of the superfluid phase due to dissipation as proposed by Bardeen Bardeen-1962 , which corresponds to the condition of the second derivative of the free energy F𝐹Fitalic_F being positive definite:

2F(q)q2=j(q)(q/m)0.superscript2𝐹𝑞superscript𝑞2𝑗𝑞𝑞𝑚0{\partial^{2}F(q)\over\partial q^{2}}={\partial j(q)\over\partial(q/m)}\geq 0\,.divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_F ( italic_q ) end_ARG start_ARG ∂ italic_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = divide start_ARG ∂ italic_j ( italic_q ) end_ARG start_ARG ∂ ( italic_q / italic_m ) end_ARG ≥ 0 . (4)

Here, the equality corresponds to the maximum (or critical) current JBsubscript𝐽𝐵J_{B}italic_J start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT that occurs at the Bardeen momentum qBsubscript𝑞𝐵q_{B}italic_q start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT. In this case, the superfuild density is defined directly as a dynamical response of the system in the form Bardeen-1962

ρs(q)j(q)(q/m),subscript𝜌𝑠𝑞𝑗𝑞𝑞𝑚\rho_{s}(q)\equiv{\partial j(q)\over\partial(q/m)}\,,italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_q ) ≡ divide start_ARG ∂ italic_j ( italic_q ) end_ARG start_ARG ∂ ( italic_q / italic_m ) end_ARG , (5)

instead of being indirectly determined in terms of the normal density (2). By this definition, ρs(q)subscript𝜌𝑠𝑞\rho_{s}(q)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_q ) vanishes at the Bardeen momentum qBsubscript𝑞𝐵q_{B}italic_q start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT where j(q)𝑗𝑞j(q)italic_j ( italic_q ) is maximum. In addition, at zero temperatures ρs(q)subscript𝜌𝑠𝑞\rho_{s}(q)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_q ) remains constant and equal to the particle density n𝑛nitalic_n when j(q)𝑗𝑞j(q)italic_j ( italic_q ) is linear in q𝑞qitalic_q for qqL𝑞subscript𝑞𝐿q\leq q_{L}italic_q ≤ italic_q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT, while it decreases to zero when qL<qqBsubscript𝑞𝐿𝑞subscript𝑞𝐵q_{L}<q\leq q_{B}italic_q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT < italic_q ≤ italic_q start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT. These considerations will be further expanded in Appendix B in terms of the free energy at the mean-field level in the presence of a current.

In the related context of superconductivity, the maximum current Jcsubscript𝐽𝑐J_{c}italic_J start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is an intrinsic property of a superconductor and corresponds to the de-pairing current, whose knowledge is crucial in applications of thin and narrow superconducting films, like nanowire single-photon detectors and microwave kinetic inductance detectors Kubo-2020 ; Kunchur-2019 ; Clem-2012 . Remaining in this related context, it may be instructive to relate the Landau threshold for viscous flow to the onset of gapless superconductivity. This can be done within the simplified context of mean-field theory, which is preparatory to the case of the BCS-BEC crossover (to be considered below) for which pairing fluctuations add to the picture. It is known since the work of Gorkov and Abrikosov Abrikosov-1960 and its generalization due to Maki Parks-1969 that a dissipative agent (like magnetic impurities, a magnetic field, or a current) can make the excitation gap in the single-particle spectrum to decrease to zero while maintaining the order parameter finite. Within mean-field (BCS) theory, when considering an imposed stationary flow with momentum 𝐪𝐪{\bf q}bold_q, the excitation spectrum is provided by the energy positions of the quasi-particle peaks in the single-particle spectral function

A(𝐤,ω;𝐪)=uk2δ(ωE+(𝐤;𝐪))+vk2δ(ω+E(𝐤;𝐪))𝐴𝐤𝜔𝐪subscriptsuperscript𝑢2𝑘𝛿𝜔subscript𝐸𝐤𝐪subscriptsuperscript𝑣2𝑘𝛿𝜔subscript𝐸𝐤𝐪A({\bf k},{\omega};{\bf q})=u^{2}_{k}\delta({\omega}-E_{+}({\bf k};{\bf q}))+v% ^{2}_{k}\delta({\omega}+E_{-}({\bf k};{\bf q}))italic_A ( bold_k , italic_ω ; bold_q ) = italic_u start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_δ ( italic_ω - italic_E start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ( bold_k ; bold_q ) ) + italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_δ ( italic_ω + italic_E start_POSTSUBSCRIPT - end_POSTSUBSCRIPT ( bold_k ; bold_q ) ) (6)

where E±(𝐤;𝐪)=E(𝐤;𝐪)±𝐤𝐪msubscript𝐸plus-or-minus𝐤𝐪plus-or-minus𝐸𝐤𝐪𝐤𝐪𝑚E_{\pm}({\bf k};{\bf q})=E({\bf k};{\bf q})\pm{{\bf k}\cdot{\bf q}\over m}italic_E start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ( bold_k ; bold_q ) = italic_E ( bold_k ; bold_q ) ± divide start_ARG bold_k ⋅ bold_q end_ARG start_ARG italic_m end_ARG and (uk2,vk2)subscriptsuperscript𝑢2𝑘subscriptsuperscript𝑣2𝑘(u^{2}_{k},v^{2}_{k})( italic_u start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) are BCS coherence factors. In this way, the total current density (1) can then be rewritten as follows

𝐣(𝐪)=n𝐪m+d𝐤(2π)3𝐤mn(𝐤;𝐪),𝐣𝐪𝑛𝐪𝑚𝑑𝐤superscript2𝜋3𝐤𝑚𝑛𝐤𝐪{\mathbf{j}}({\bf q})=n{{\bf q}\over m}+\!\int\!{d{\bf k}\over(2\pi)^{3}}\;{{% \bf k}\over m}\;n({\bf k};{\bf q}),bold_j ( bold_q ) = italic_n divide start_ARG bold_q end_ARG start_ARG italic_m end_ARG + ∫ divide start_ARG italic_d bold_k end_ARG start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG divide start_ARG bold_k end_ARG start_ARG italic_m end_ARG italic_n ( bold_k ; bold_q ) , (7)

where we have introduced the density distribution function

n(𝐤;𝐪)=2+𝑑ωf(ω)A(𝐤,ω;𝐪).𝑛𝐤𝐪2superscriptsubscriptdifferential-d𝜔𝑓𝜔𝐴𝐤𝜔𝐪n({\bf k};{\bf q})=2\int_{-\infty}^{+\infty}\!\!\!d{\omega}f({\omega})A({\bf k% },{\omega};{\bf q}).italic_n ( bold_k ; bold_q ) = 2 ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + ∞ end_POSTSUPERSCRIPT italic_d italic_ω italic_f ( italic_ω ) italic_A ( bold_k , italic_ω ; bold_q ) . (8)

In the simplest case of T0𝑇0T\rightarrow 0italic_T → 0 when thermal dissipation is absent, the onset of viscous flow given by the Landau criterion

E(𝐤,𝐪)kqm=0𝐸𝐤𝐪𝑘𝑞𝑚0E({\bf k},{\bf q})-{kq\over m}=0italic_E ( bold_k , bold_q ) - divide start_ARG italic_k italic_q end_ARG start_ARG italic_m end_ARG = 0 (9)

implies that the position of one of the two peaks in the spectral function (6) has shifted to zero energy, thereby signaling the closure of the excitation gap at some value of 𝐤𝐤{\bf k}bold_k while the order parameter Δ𝐪subscriptΔ𝐪\Delta_{{\bf q}}roman_Δ start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT remains finite. As a consequence, the Landau threshold at qLsubscript𝑞𝐿q_{L}italic_q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT in the zero-temperature curve of Fig.1 implies the closing of the excitation gap in the single-particle spectral function (or, more conventionally, in the density of states Fulde-1965 ; Parks-1969 ). This discussion will be relevant in the next Section when discussing the inclusion of pairing fluctuations.

Refer to caption
Figure 2: Momentum dependence of the current j(q)𝑗𝑞j(q)italic_j ( italic_q ) for a weakly-repulsive Bose gas at several temperatures in the superfluid phase. At finite temperature, a maximum develops corresponding to the critical current Jcsubscript𝐽𝑐J_{c}italic_J start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, which as T0𝑇0T\rightarrow 0italic_T → 0 coincides with the current flowing at the sound velocity. The inset shows a comparison between the energy spectra of the Bose gas at rest and moving at the sound speed.

II.2 Intrinsic critical current of a
       a weakly-repulsive Bose gas

The current density response of a weakly-interacting Bose gas moving with uniform momentum 𝐪𝐪{\bf q}bold_q can similarly be studied Fetter-1972 . In this case, the current density has an expression similar to Eq. (1) with suitable replacements, namely,

𝐣(𝐪)=n𝐪M+d𝐐(2π)3𝐐Mb(E+B(𝐐;𝐪))𝐣𝐪𝑛𝐪𝑀𝑑𝐐superscript2𝜋3𝐐𝑀𝑏superscriptsubscript𝐸𝐵𝐐𝐪{\mathbf{j}}({\bf q})=n{{\bf q}\over M}+\int\!{d{\bf Q}\over(2\pi)^{3}}\;{{\bf Q% }\over M}\;b(E_{+}^{B}({\bf Q};{\bf q}))bold_j ( bold_q ) = italic_n divide start_ARG bold_q end_ARG start_ARG italic_M end_ARG + ∫ divide start_ARG italic_d bold_Q end_ARG start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG divide start_ARG bold_Q end_ARG start_ARG italic_M end_ARG italic_b ( italic_E start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( bold_Q ; bold_q ) ) (10)

where n𝑛nitalic_n is the (bosonic) particle density, M𝑀Mitalic_M the boson mass, E+B(𝐐;𝐪)=EB(𝐐,q)+𝐐𝐪Msuperscriptsubscript𝐸𝐵𝐐𝐪superscript𝐸𝐵𝐐𝑞𝐐𝐪𝑀E_{+}^{B}({\bf Q};{\bf q})=E^{B}({\bf Q},q)+{{\bf Q}\cdot{\bf q}\over M}italic_E start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( bold_Q ; bold_q ) = italic_E start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( bold_Q , italic_q ) + divide start_ARG bold_Q ⋅ bold_q end_ARG start_ARG italic_M end_ARG the Doppler-shifted Bogoliubov quasi-particle spectrum with EB(𝐐,q)=(𝐐22M+n0(T,q)g)2(n0(T,q)g)2superscript𝐸𝐵𝐐𝑞superscriptsuperscript𝐐22𝑀subscript𝑛0𝑇𝑞𝑔2superscriptsubscript𝑛0𝑇𝑞𝑔2E^{B}({\bf Q},q)=\sqrt{\left({{\bf Q}^{2}\over 2M}+n_{0}(T,q)g\right)^{2}-% \left(n_{0}(T,q)g\right)^{2}}italic_E start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( bold_Q , italic_q ) = square-root start_ARG ( divide start_ARG bold_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_M end_ARG + italic_n start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_T , italic_q ) italic_g ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - ( italic_n start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_T , italic_q ) italic_g ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG, g𝑔gitalic_g the coupling constant, n0(T,q)subscript𝑛0𝑇𝑞n_{0}(T,q)italic_n start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_T , italic_q ) the condensate density as determined consistently through the density equation for each imposed momentum q, and b(ϵ)=(eϵkBT1)1𝑏italic-ϵsuperscriptsuperscript𝑒italic-ϵsubscript𝑘𝐵𝑇11b(\epsilon)=(e^{\epsilon\over k_{B}T}-1)^{-1}italic_b ( italic_ϵ ) = ( italic_e start_POSTSUPERSCRIPT divide start_ARG italic_ϵ end_ARG start_ARG italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG end_POSTSUPERSCRIPT - 1 ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT is the Bose distribution function.

In Fig. 2 the induced current is shown as a function of the impressed momentum 𝐪𝐪{\bf q}bold_q for several temperatures below the critical temperature Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT (which coincides with that of the non-interacting Bose gas, as consistently determined by the Bogoliubov theory), with the sound velocity cs=n0(T=0,q=0)gMsubscript𝑐𝑠subscript𝑛0formulae-sequence𝑇0𝑞0𝑔𝑀c_{s}=\sqrt{n_{0}(T=0,q=0)g\over M}italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = square-root start_ARG divide start_ARG italic_n start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_T = 0 , italic_q = 0 ) italic_g end_ARG start_ARG italic_M end_ARG end_ARG being used for normalization in both axes. At zero temperature, the straight line sharply drops to -\infty- ∞ as soon as the velocity exceeds cssubscript𝑐𝑠c_{s}italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, thus signaling an instability of the system. At finite temperature, the instability limit corresponds to the maximum of each curve, and represents the maximum (critical) current Jcsubscript𝐽𝑐J_{c}italic_J start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT that can be sustained before a destructive flow sets in.

Beyond this point (and depending on the nature of the excitation spectrum of the Bose gas) a number of inhomogeneous phases have been considered in the literature, ranging from the condensation of rotons Iorda-1980 to the Bogoliubov-Cerenkov radiation Caru-2006 ; Baym-2012 , as well as to the recent observation of rotonic density modulations and supersolidity Modugno-2019 . In the present work, we limit ourselves quite generally to considering the behavior of the current up to but not past the critical current.

III Intrinsic critical current
throughout the BCS-BEC crossover

In this Section, we perform a comprehensive study of the current-vs-momentum characteristics in the presence of pairing fluctuations for a fermionic superfluid spanning the BCS-BEC crossover. By following the onset of dissipation signaled by these characteristics, we will show how the Landau critical velocity can be identified for all couplings from the BCS to BEC limits. This will be done by explicitly calculating a single measurable quantity, rather than by merely identifying the branches of the single-particle and two-particle excitation spectra corresponding to the lower velocity for given coupling, as one would instead do in a BCS-RPA approach Comb-2006 ; Spunta-2010 .

III.1 Current-vs-momentum characteristics
       in the presence of pairing fluctuations

The fermionic expression (7) for the total current density can quite generally be generalized to include pairing fluctuations, by adopting the following definition for the density distribution function (cf. Ref. Pisani-2023 and Appendix A below):

n(𝐤;𝐪)=2kBTneiωnη𝒢11pf(𝐤,ωn;𝐪).𝑛𝐤𝐪2subscript𝑘𝐵𝑇subscript𝑛superscript𝑒𝑖subscript𝜔𝑛𝜂superscriptsubscript𝒢11pf𝐤subscript𝜔𝑛𝐪n({\bf k};{\bf q})=2k_{B}T\sum_{n}e^{i{\omega}_{n}\eta}\,{\cal G}_{11}^{% \mathrm{pf}}({\bf k},{\omega}_{n};{\bf q})\,.italic_n ( bold_k ; bold_q ) = 2 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T ∑ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_η end_POSTSUPERSCRIPT caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_pf end_POSTSUPERSCRIPT ( bold_k , italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ; bold_q ) . (11)

Here, ωn=(2n+1)πkBTsubscript𝜔𝑛2𝑛1𝜋subscript𝑘𝐵𝑇{\omega}_{n}=(2n+1)\pi k_{B}Titalic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = ( 2 italic_n + 1 ) italic_π italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T (n𝑛nitalic_n integer) is a fermionic Matsubara frequency, η𝜂\etaitalic_η a positive infinitesimal, and 𝒢11pfsuperscriptsubscript𝒢11pf{\cal G}_{11}^{\mathrm{pf}}caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_pf end_POSTSUPERSCRIPT the “normal” single-particle Green’s function obtained in the presence of a superfluid flow with momentum 𝐪𝐪{\bf q}bold_q. In the following, the single-particle Green’s function 𝒢11pfsuperscriptsubscript𝒢11pf{\cal G}_{11}^{\mathrm{pf}}caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_pf end_POSTSUPERSCRIPT will be calculated within the t𝑡titalic_t-matrix approach in the presence of a super-current as developed in Ref. Pisani-2023 (although when comparing with experimental data the refinements provided by the extended GMB approach of Ref. Pisani-2018-b will also be considered). Like in the mean-field case (cf. Eq. (1)), the current density given by Eqs. (7) and (11) is again made up of two terms, a standard (classical) term proportional to the carriers velocity and a dissipative term identified as a back-flow current.

Refer to caption
Figure 3: The current-vs-momentum characteristics (normalized to JF=kFn/msubscript𝐽𝐹subscript𝑘𝐹𝑛𝑚J_{F}=k_{F}n/mitalic_J start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_n / italic_m) for three representative couplings across the BCS-BEC crossover and several temperatures in the superfluid phase, as obtained by the t𝑡titalic_t-matrix approach, are shown in the left panels. For given coupling and temperature, the value of the critical current corresponds to the maximum of the curve according to the Bardeen criterion (where the temperature is now in units of the superfluid critical temperature Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT calculated with the inclusion of pairing fluctuations). At zero temperature, both Landau (JLsubscript𝐽𝐿J_{L}italic_J start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT) and Bardeen (JBsubscript𝐽𝐵J_{B}italic_J start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT) critical currents can be identified, as shown at unitarity with their identifying labels in the left-central panel. In addition, the right panels report the temperature dependence of the Bardeen critical current for given coupling.

Figure 3 shows the current versus momentum characteristics obtained from Eqs. (7) and (11), for three representative couplings across the BCS-BEC crossover and several temperatures (left panels). At low-enough temperatures, the behavior is analogous to the mean-field case of Sec. II.1. Accordingly, a threshold is found at a critical momentum, past which the current density is seen to deviate from a linear behavior and to enter a dissipative regime due to the onset of the back-flow current. Below we will argue that this critical value corresponds to the Landau critical velocity qL/msubscript𝑞𝐿𝑚q_{L}/mitalic_q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT / italic_m, where now the single-particle excitation branch of the spectrum is suitably renormalized by pairing fluctuations.

Similarly to the mean-field case of Sec. II.1, at any temperature the current density is seen to saturate at a maximum value, labelled in Fig. 3 by JBsubscript𝐽𝐵J_{B}italic_J start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT in correspondence to the Bardeen momentum qBsubscript𝑞𝐵q_{B}italic_q start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT. This value corresponds to the Bardeen criterion (4) for the thermodynamic stability of the intrinsic current, such that past qBsubscript𝑞𝐵q_{B}italic_q start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT the system enters into a new phase. The temperature dependence of the intrinsic critical current JBsubscript𝐽𝐵J_{B}italic_J start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT for various couplings is reported in the right panels of Fig. 3. Note how the behavior of this temperature dependence confirms what was recently found in Ref. Pisani-2023 (cf. Fig. 5 therein) for the critical current that can flow through a barrier, whose temperature dependence changes from a convex to a concave behavior from the BCS to the BEC regime, passing through an essentially linear behavior at unitarity. This linear temperature behavior of the critical current of a strongly-interacting fermionic superfluid is reminiscent of an analogous behavior of the critical velocity of 44{}^{4}start_FLOATSUPERSCRIPT 4 end_FLOATSUPERSCRIPTHe (as shown in Fig. 12 of Ref. Varoquaux-2015 ).

Note also from Fig. 3 that the maximum of the characteristics about qBsubscript𝑞𝐵q_{B}italic_q start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT is much broader at unitarity than on either side of the crossover, and that the values of JLsubscript𝐽𝐿J_{L}italic_J start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT and JBsubscript𝐽𝐵J_{B}italic_J start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT remain close to each other in contrast to their respective momenta qLsubscript𝑞𝐿q_{L}italic_q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT and qBsubscript𝑞𝐵q_{B}italic_q start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT which differ appreciably from each other. Here, a clear (albeit not sharp) onset of dissipation is not restricted to zero temperature but extends, in practice, up to about 0.5Tc÷0.6Tc0.5subscript𝑇𝑐0.6𝑇𝑐0.5T_{c}\div 0.6T{c}0.5 italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ÷ 0.6 italic_T italic_c, similarly to what was found in Sec. II.1 for a weakly-attractive Fermi gas. This result is consistent with the fact that, in the broken-symmetry phase, the Landau damping due to thermal fluctuations becomes relevant above 0.3Tc÷0.4Tc0.3subscript𝑇𝑐0.4subscript𝑇𝑐0.3T_{c}\div 0.4T_{c}0.3 italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ÷ 0.4 italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, although the quasi-particle approximation (or, equivalently, the presence of a sharp energy spectrum on which the Landau criterion rests to begin with) ceases to be valid only above 0.5÷0.6Tc0.50.6subscript𝑇𝑐0.5\div 0.6T_{c}0.5 ÷ 0.6 italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT Pieri-2004 . We shall return to this point below.

III.2 Landau and Bardeen intrinsic critical velocities at low temperature

Quite generally, (quantum and thermal) pairing fluctuations are expected to give rise to a finite lifetime in the energy excitations, making the quasi-particle approximation much less reliable than in the mean-field case of Sec. II.1. Nevertheless, by analyzing in detail the single- and two-particle excitation spectra in the presence of pairing fluctuations throughout the BCS-BEC crossover, in Ref. Pieri-2004 it was shown that at low-enough temperature a quasi-particle picture is still applicable to the extent that the energy spectrum remains rather sharp. In addition, it was shown that the single-particle spectral function A(k,ω)𝐴𝑘𝜔A(k,{\omega})italic_A ( italic_k , italic_ω ) possesses BCS-like features, with coherent peaks of quite small line-widths (cf. Fig. 10 of Ref. Pieri-2004 ) and dispersions showing a characteristic back-bending (cf. Fig. 13 of Ref. Pieri-2004 ). It was further shown that the degeneracy between the order parameter and the excitation gap occurring at the mean-field level is removed by pairing fluctuations. Specifically, a quantitative comparison between these two energy scales at low temperature was reported in Fig. 14 of Ref. Pieri-2004 , showing a reduction of about 10%percent1010\%10 % of the excitation gap with respect to the order parameter at unitarity.

Accordingly, as long as at low temperature the single-particle spectrum can be interpreted in terms of quasi-particle excitations with well-defined coherent peaks, one may assume a BCS-like expression like that given by Eq. (6) for the single-particle spectral function to be valid throughout the BCS-BEC crossover (although with renormalized values of the thermodynamic parameters appearing therein), thereby taking advantage of the arguments discussed in Sec. II.1. This would imply that, even in the presence of pairing fluctuations, the onset of the non-linear behavior in 𝐣(𝐪)𝐣𝐪{\mathbf{j}}({\bf q})bold_j ( bold_q ) should reflect the closing of the excitation gap, thereby identifying the Landau critical threshold.

Refer to caption
Figure 4: Coupling dependence of the critical velocity qc/msubscript𝑞𝑐𝑚q_{c}/mitalic_q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_m at low temperature extracted from the characteristics j(q)𝑗𝑞j(q)italic_j ( italic_q ) for various couplings. Here, the red solid line represents the Landau velocity for which qc=qLsubscript𝑞𝑐subscript𝑞𝐿q_{c}=q_{L}italic_q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = italic_q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT and the green dashed line the Bardeen velocity for which qc=qBsubscript𝑞𝑐subscript𝑞𝐵q_{c}=q_{B}italic_q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = italic_q start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT, both obtained like in Fig. 3 by the t𝑡titalic_t-matrix approach. The blue dotted line corresponds to the critical velocity for pair-breaking excitations as obtained from the spectral function. Long-dashed lines (dashed-dotted lines) correspond to pair-breaking (ascending branch) and sound (descending branch) critical velocities, with thermodynamic parameters that do not include (include) pairing fluctuations Comb-2006 ; Spunta-2010 (see the text).

Figure 4 reports several curves for the critical velocity vc=qc/msubscript𝑣𝑐subscript𝑞𝑐𝑚v_{c}=q_{c}/mitalic_v start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = italic_q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_m throughout the BCS-BEC crossover at low temperature, with qcsubscript𝑞𝑐q_{c}italic_q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT obtained from different approximations. All curves contain an “ascending” branch associated with pair-breaking excitations in the BCS side of the crossover and a “descending” branch associated with phonon excitations in the BEC side of the crossover Spunta-2010 .

Initially, a BCS-RPA approach Anderson-1958 is adopted and extended throughout the BCS-BEC crossover MPS-1998 ; Spunta-2010 ; Comb-2006 , whereby the pair-breaking branch is obtained via the expression (3) for the Landau momentum qLsubscript𝑞𝐿q_{L}italic_q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT and the phonon branch via the expression of the sound velocity at zero temperature reported in Ref. MPS-1998 , with the thermodynamic parameters ΔΔ\Deltaroman_Δ and μ𝜇\muitalic_μ calculated either at the mean-field level (long-dashed lines Spunta-2010 ; Comb-2006 ) or with the further inclusion of pairing fluctuations (dashed-dotted lines). Both these pairs of lines represent an upper bound to the Landau critical velocity (cf. Fig. 8 of Ref. Comb-2006 and Fig. 24 of Ref. Spunta-2010 ), and become strongly renormalized by dynamical many-body effects once properly included.

In this respect, the red solid line represents the value of the Landau critical velocity vL=qL/msubscript𝑣𝐿subscript𝑞𝐿𝑚v_{L}=q_{L}/mitalic_v start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT = italic_q start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT / italic_m obtained at a low (but non-zero) temperature (T=0.1Tc𝑐{}_{c}start_FLOATSUBSCRIPT italic_c end_FLOATSUBSCRIPT) from the threshold of the non-linear behavior as identified in Fig. 3. Note here the strong suppression with respect to the previous BCS-RPA results (black long-dashed Comb-2006 ; Spunta-2010 and dashed-dotted lines), which is due to dynamical effects that are not taken into account in those results, yielding a significant renormalization of the excitation gap and of the underlying Fermi surface. In addition, the short-dashed line of Fig. 4 represents the velocity qB/msubscript𝑞𝐵𝑚q_{B}/mitalic_q start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT / italic_m for which the current reaches its critical value, corresponding to the maximum value extracted from the left panels of Fig. 3.

By taking advantage of the analytic considerations reported in Sec. II.1, the single-particle excitations at low temperature can be interpreted throughout the crossover in terms of a quasi-particle approximation with an effective BCS-like single-particle spectral function. Accordingly, following Ref. Pieri-2004 , for given coupling we have obtained renormalized values of the order parameter and chemical potential by a BCS-like fit to the dispersion relation extracted from the single-particle spectral function, in terms of which we have calculated the Landau critical velocity for the ascending branch using the expression (3). The result of this calculation corresponds to the (blue) dotted line of Fig. 4. Note how this line is quite close although not identical to the solid line therein, owing to a small broadening of the spectral line-shape associated with quantum fluctuations not captured by the quasi-particle approximation.

Physically, the dotted line of Fig. 4 represents the Landau velocity relative to the single-particle (pair-breaking) contribution to the excitation spectrum. This is the reason why it tends to increase without bound upon approaching the BEC regime, where it is expected to follow the behavior of the pair binding energy. The solid line of Fig. 4 instead switches smoothly from the ascending branch of pair-breaking excitations to the descending branch of phonon excitations, with the maximum reached for coupling (kFaF)1=+0.36superscriptsubscript𝑘𝐹subscript𝑎𝐹10.36(k_{F}a_{F})^{-1}=+0.36( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = + 0.36 (cf. Appendix B for for further considerations on this topic). This smooth evolution of the intrinsic critical velocity, from the ascending to the descending branches when the BCS-BEC crossover is spanned from the BCS to the BEC regimes, takes place directly from the expressions (7) and (11) for the homogeneous case where pairing fluctuations are included. It is worth pointing out that this smooth evolution corresponds to a more realistic picture of the critical velocity, by removing the unphysical cusp present in the BCS-RPA approach (long-dashed and dashed-dotted black lines in Fig. 4) Spunta-2010 ; Comb-2006 . At the mean-field level, on the other hand, a smooth evolution between the two branches can be recovered when the current is made to flow in the presence of a small (in the limit, infinitesimal) barrier that breaks translational invariance Spunta-2010 ; Piselli-2020 .

Refer to caption
Figure 5: Comparison with the experimental data for the Landau critical velocity at low temperature from Ref. Miller-2007 (light green asterisk) and Ref. Weimer-2015 (dark green squares). Theoretical results obtained by the extended GMB approach of Ref. Pisani-2018-b in the homogeneous case are given by the blue solid line, and are compared with the results obtained by the t𝑡titalic_t-matrix approach reproduced from Fig. 4 above (red dashed-dotted line). Results obtained with the addition of a small Gaussian barrier (as specified in Ref. Weimer-2015 for different number of atoms) are given by light and dark circles (joined by dotted lines for clarity) The corresponding results obtained by the Gross-Pitaevskii equation for composite bosons with scattering length 0.6aF0.6subscript𝑎𝐹0.6a_{F}0.6 italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT are reported as light and dark diamonds. For completeness, the long-dashed line reproduces the BCS-RPA critical velocity also reported in Fig. 4 above.

In this respect, it may be instructive to consider the BEC regime of Eqs. (7) and (11), for which analytic expressions can be obtained (cf. Appendix B of Ref. Pisani-2023 ). One obtains the following approximate expression for the density distribution function (11):

n(𝐤,𝐪)14ξ(𝐤)2[2Δ𝐪2+2d𝐐(2π)3kBTνeiΩνηΓ11(𝐐,Ων;𝐪)(1+𝐤𝐐/mξ(𝐤))]similar-to-or-equals𝑛𝐤𝐪14𝜉superscript𝐤2delimited-[]2superscriptsubscriptΔ𝐪22𝑑𝐐superscript2𝜋3subscript𝑘𝐵𝑇subscript𝜈superscript𝑒𝑖subscriptΩ𝜈𝜂subscriptΓ11𝐐subscriptΩ𝜈𝐪1𝐤𝐐𝑚𝜉𝐤n({\bf k},{\bf q})\simeq\frac{1}{4\xi(\mathbf{k})^{2}}\left[2\Delta_{\bf q}^{2% }\;+\right.\\ \left.2\!\!\int\!\!\!\frac{d\mathbf{Q}}{(2\pi)^{3}}k_{B}T\sum_{\nu}\!e^{i% \Omega_{\nu}\eta}\,\Gamma_{11}(\mathbf{Q},\Omega_{\nu};\mathbf{q})\left(1+% \frac{\mathbf{k}\cdot\mathbf{Q}/m}{\xi(\mathbf{k})}\right)\right]start_ROW start_CELL italic_n ( bold_k , bold_q ) ≃ divide start_ARG 1 end_ARG start_ARG 4 italic_ξ ( bold_k ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG [ 2 roman_Δ start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + end_CELL end_ROW start_ROW start_CELL 2 ∫ divide start_ARG italic_d bold_Q end_ARG start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T ∑ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i roman_Ω start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT italic_η end_POSTSUPERSCRIPT roman_Γ start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( bold_Q , roman_Ω start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ; bold_q ) ( 1 + divide start_ARG bold_k ⋅ bold_Q / italic_m end_ARG start_ARG italic_ξ ( bold_k ) end_ARG ) ] end_CELL end_ROW (12)

where ξ(𝐤)=𝐤2/(2m)μ𝜉𝐤superscript𝐤22𝑚𝜇\xi(\mathbf{k})=\mathbf{k}^{2}/(2m)-\muitalic_ξ ( bold_k ) = bold_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 2 italic_m ) - italic_μ, Ων=2πνkBTsubscriptΩ𝜈2𝜋𝜈subscript𝑘𝐵𝑇\Omega_{\nu}=2\pi\nu k_{B}Troman_Ω start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT = 2 italic_π italic_ν italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T (ν𝜈\nuitalic_ν integer) is a bosonic Matsubara frequency, and Γ11(𝐐,Ων;𝐪)subscriptΓ11𝐐subscriptΩ𝜈𝐪\Gamma_{11}(\mathbf{Q},\Omega_{\nu};\mathbf{q})roman_Γ start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( bold_Q , roman_Ω start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ; bold_q ) is the diagonal element of the particle-particle ladder in the presence of a super-current Pisani-2023 . After integration over the fermionic variables, the total current density (7) becomes Pisani-2023 :

𝐣(𝐪)n𝐪m2d𝐐(2π)3𝐐2mb(E+B(𝐐;𝐪))similar-to-or-equals𝐣𝐪𝑛𝐪𝑚2𝑑𝐐superscript2𝜋3𝐐2𝑚𝑏superscriptsubscript𝐸𝐵𝐐𝐪{\mathbf{j}}({\bf q})-n{{\bf q}\over m}\simeq 2\!\int\!\frac{d\mathbf{Q}}{(2% \pi)^{3}}\,\frac{\mathbf{Q}}{2m}\,b\!\left(E_{+}^{B}(\mathbf{Q};\mathbf{q})\right)bold_j ( bold_q ) - italic_n divide start_ARG bold_q end_ARG start_ARG italic_m end_ARG ≃ 2 ∫ divide start_ARG italic_d bold_Q end_ARG start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG divide start_ARG bold_Q end_ARG start_ARG 2 italic_m end_ARG italic_b ( italic_E start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( bold_Q ; bold_q ) ) (13)

where E+B(𝐐)superscriptsubscript𝐸𝐵𝐐E_{+}^{B}(\mathbf{Q})italic_E start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( bold_Q ) is given after Eq. (10) and 2m=M2𝑚𝑀2m=M2 italic_m = italic_M is the mass of composite bosons that form in the BEC limit (to which an imposed momentum 2𝐪2𝐪2\mathbf{q}2 bold_q is now associated). It is thus evident that the fermionic current has become purely bosonic in nature, with the Landau threshold now set by the Bogoliubov sound velocity (cf. Sec. II.2).

III.3 Comparison with the available experimental data for the critical velocity

A direct comparison can be made at this point with the available experimental data for the Landau critical velocity vLsubscript𝑣𝐿v_{L}italic_v start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT obtained with an ultra-cold trapped Fermi gas, initially at unitarity in Ref. Miller-2007 and, more extensively, throughout the BCS-BEC crossover in Ref. Weimer-2015 . These data are reported in Fig. 5 as a light green asterisk Miller-2007 and dark green squares Weimer-2015 , respectively. To make the best possible comparison with these experimental data, we have improved on the t𝑡titalic_t-matrix results shown previously in Fig. 4 and implemented in the present context the extended GMB approach of Ref. Pisani-2018-b , which has recently proved to lead to a quite good comparison with experiments in several contexts Moritz-2022 ,Koehl-2023 ,Pisani-2023 ,Piselli-2023 . [For the benefit of the readers, a concise summary of the extended GMB approach of Ref. Pisani-2018-b can be found in Appendix A below.]

The results of the extended GMB approach for the critical Landau velocity at low temperature in the homogeneous case are shown by the blue solid line of Fig. 5.

To offer a direct comparison with the experimental results, Fig. 5 also reproduces from Fig. 4 both the results obtained by the t𝑡titalic_t-matrix approach (red dashed-dotted line) and the BCS-RPA critical velocity (long-dashed line).

A quite good agreement is found between the experimental data and the extended GMB approach for the ascending (pair-breaking) branch, at least for coupling values up to about +0.20.2+0.2+ 0.2 before the descending (phononic) branch prevails and the maximum occurs, with a slight (yet favorable) improvement on the BCS side of the crossover over and above the results of the t𝑡titalic_t-matrix approach. It turns out, however, that the behavior of vLsubscript𝑣𝐿v_{L}italic_v start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT about and past this (intermediate) coupling regime is sensible to the presence even of the small barrier considered in the experiment of Ref. Weimer-2015 . To quantify this effect, on top of the extended GMB approach we have considered a small Gaussian barrier like that utilized in the experiment of Ref. Weimer-2015 (although of a different sign, which should be immaterial in the limit of infinitesimal barrier), with the results given by the light and dark circles in Fig. 5.

Here, the presence of this barrier is dealt with the mLPDA approach of Ref. Pisani-2018-b , where pairing fluctuations are included on top of the original LPDA approach of Ref. Simonucci-2014 . [For the benefit of the readers, a concise summary of the LPDA approach of Ref. Simonucci-2014 as well as of the mLPDA approach of Ref. Pisani-2018-b can be found in Appendix A below.]

The difference between these values stems from the experimental uncertainty in the number of atoms Weimer-2015 , which is reflected in the values of the Fermi momentum utilized to normalize the theoretical results (see below).

It is evident from Fig. 5 that the presence of a small barrier has only minor effects on the ascending branch on the BCS side of the crossover up to (about) unitarity, but its effects become rather substantial on the BEC side of the crossover past unitarity when the descending branch is dominated by bosonic degrees of freedom and the underlying Fermi surface is lost. This enhanced sensitivity to the presence of spatial inhomogeneities on the BEC side of the crossover is in line with what found in Ref. Palest-2013 for the effects of random impurities.

Yet, in Fig. 5 discrepancies with the experimental data still remain deep in the BEC side of the crossover for (kFaF)11superscriptsubscript𝑘𝐹subscript𝑎𝐹11(k_{F}a_{F})^{-1}\approx 1( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ≈ 1, even after having included the effects of a small barrier. This discrepancy could be due to our theory recovering in this limit the (Born) value 2.0aF2.0subscript𝑎𝐹2.0a_{F}2.0 italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT Pieri-2003 instead of the correct value 0.6aF0.6subscript𝑎𝐹0.6a_{F}0.6 italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT Brodsky-2006 for the scattering length of composite bosons, thereby overestimating the value of the speed of sound. To clarify this point, we have calculated the Landau critical velocity at zero temperature for coupling (kFaF)1=1.0superscriptsubscript𝑘𝐹subscript𝑎𝐹11.0(k_{F}a_{F})^{-1}=1.0( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = 1.0 in terms of the Gross-Pitaevskii equation, with the values 0.6aF0.6subscript𝑎𝐹0.6a_{F}0.6 italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT for the bosonic scattering length. The results of this additional calculation are shown in Fig. 5 by light and dark diamonds, respectively, where for consistency the presence of a small barrier is also taken into account as we did above. One sees that, with these additions, in the BEC regime the theoretical results come considerably closer to the experimental data.

The residual discrepancy between theory and experiment, as far as the bosonic side of the descending branch in Fig. 5 is concerned, can be ascribed to the method used in Ref. Weimer-2015 to excite the gas of ultra-cold atoms through a circular stirring of a laser. The same authors of Ref. Weimer-2015 have proven in Ref. Weimer-2016 that for a bosonic gas the additional centrifugal energy present in a circular stirring acts to lower significantly the value of the Landau critical velocity, which would instead be expected to coincide with the Bogoliubov sound velocity (see also Fig. 2(a) in Ref. Weimer-2015 ). In contrast, the presence of an underlying Fermi surface on the BCS side of the crossover up to somewhat past unitarity makes the centrifugal energy negligible when compared with the fermionic chemical potential.

We return, finally, to the values of the Fermi momentum mentioned above and utilized to normalize the theoretical results. In Ref. Miller-2007 , the experimental values of vLsubscript𝑣𝐿v_{L}italic_v start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT were expressed directly in terms of the local Fermi velocity vFlocsuperscriptsubscript𝑣𝐹locv_{F}^{\mathrm{loc}}italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_loc end_POSTSUPERSCRIPT at the trap center, yielding the value vL/vFloc=0.25subscript𝑣𝐿superscriptsubscript𝑣𝐹loc0.25v_{L}/v_{F}^{\mathrm{loc}}=0.25italic_v start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT / italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_loc end_POSTSUPERSCRIPT = 0.25 reported in Fig. 5 (light green asterisk). In Ref. Weimer-2015 , on the other hand, the experimental values of vLsubscript𝑣𝐿v_{L}italic_v start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT were given in terms of the global (trap) Fermi velocity vFtsuperscriptsubscript𝑣𝐹𝑡v_{F}^{t}italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT. To estimate the corresponding local values of vFlocsuperscriptsubscript𝑣𝐹locv_{F}^{\mathrm{loc}}italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_loc end_POSTSUPERSCRIPT at the trap center, we have taken advantage of the procedure followed in Ref. Weimer-2015 when converting the theoretical values for the speed of sound obtained in Ref. Astra-2004 in the homogeneous case to the experimental trap geometry throughout the whole BCS-BEC crossover. This procedure effectively amounts to multiplying the value of the sound velocity obtained in the homogeneous case by the factor vFloc/vFtsuperscriptsubscript𝑣𝐹locsuperscriptsubscript𝑣𝐹𝑡v_{F}^{\mathrm{loc}}/v_{F}^{t}italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_loc end_POSTSUPERSCRIPT / italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT. We have used this procedure in reverse and extracted the geometrical factor vFt/vFlocsuperscriptsubscript𝑣𝐹𝑡superscriptsubscript𝑣𝐹locv_{F}^{t}/v_{F}^{\mathrm{loc}}italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT / italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_loc end_POSTSUPERSCRIPT specific to the trap settings of Ref. Weimer-2015 , which has enabled us to convert the value of the experimental data of Ref. Weimer-2015 for the Landau critical velocity in a way to compare them with our theoretical results for the homogeneous case.

III.4 Josephson critical current interpreted as
       an intrinsic critical current inside the barrier

Refer to caption
Figure 6: The Josephson characteristic, obtained at unitarity and T/Tc=0.15𝑇subscript𝑇𝑐0.15T/T_{c}=0.15italic_T / italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 0.15 for a Gaussian barrier of width 2.5kF12.5superscriptsubscript𝑘𝐹12.5k_{F}^{-1}2.5 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and height V0/EF=0.1subscript𝑉0subscript𝐸𝐹0.1V_{0}/E_{F}=0.1italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = 0.1, as expressed in terms of the local superfluid velocity vs(x=0)subscript𝑣𝑠𝑥0v_{s}(x=0)italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_x = 0 ) at the barrier center (filled squares), is compared with the intrinsic current obtained for a homogeneous system using the “local” thermodynamic parameters determined at the barrier center (filled circles).

The considerations made so far for the intrinsic critical current (or critical velocity) have naturally direct applications to physical systems which are either homogeneous or slightly deviate from a homogeneous condition. In this Section we apply the concept of critical current to the case of the Josephson effect for a fermionic superfluid flowing across a barrier of width larger than or comparable with the healing length of the bulk superfluid. This condition is satisfied by most experiments with ultra-cold Fermi atomic gases, for which the healing length is of the order of the inter-particle spacing kF1superscriptsubscript𝑘𝐹1k_{F}^{-1}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT in the relevant coupling range 0.5<(kFaF)1<10.5superscriptsubscript𝑘𝐹subscript𝑎𝐹11-0.5<(k_{F}a_{F})^{-1}<1- 0.5 < ( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT < 1 probed by the experiments, while the width of the barrier cannot be made as small Kwon-2020 ; DelPace-2021 .

In the Josephson effect, the maximum (critical) current that can flow across a barrier depends on the shape (height and width) of the barrier, besides coupling and temperature. In Ref. Pisani-2023 a systematic study of the Josephson characteristics was performed in terms of a theoretical (mLPDA) approach that allows for the inclusion of pairing fluctuations in a non-trivial spatial geometry, thus complementing and extending the original LPDA approach of Ref. Simonucci-2014 . [For the benefit of the readers, a concise summary of the LPDA approach of Ref. Simonucci-2014 as well as of the mLPDA approach of Ref. Pisani-2018-b can be found in Appendix A below.]

Here, we analyze the case examined in Figs. 2 and 3 of Ref. Pisani-2023 , where a Gaussian barrier of width 2.5kF12.5superscriptsubscript𝑘𝐹12.5k_{F}^{-1}2.5 italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and height V0/EF=0.1subscript𝑉0subscript𝐸𝐹0.1V_{0}/E_{F}=0.1italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = 0.1 was considered for temperature T/Tc=0.15𝑇subscript𝑇𝑐0.15T/T_{c}=0.15italic_T / italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 0.15 at unitarity. It turns out that the maximum current sustained by the barrier is quantitatively related to the intrinsic critical current of a bulk superfluid with the same “local” thermodynamic conditions that develop at the center of the barrier.

This behavior is shown in Fig. 6, where the current density as obtained by the methods of Ref. Pisani-2023 is reported as a function of the local superfluid velocity vs(x=0)subscript𝑣𝑠𝑥0v_{s}(x=0)italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_x = 0 ) at the center of the barrier, and compared with the intrinsic current 𝐣(𝐪)𝐣𝐪{\mathbf{j}}({\bf q})bold_j ( bold_q ) where |𝐪|=mvs𝐪𝑚subscript𝑣𝑠|{\bf q}|=mv_{s}| bold_q | = italic_m italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT as obtained from Eqs. (7) and (11) for a homogeneous superfluid with the same coupling and temperature, in which the local thermodynamic parameters Δ(x=0)Δ𝑥0\Delta(x=0)roman_Δ ( italic_x = 0 ), μ(x=0)𝜇𝑥0\mu(x=0)italic_μ ( italic_x = 0 ), and n(x=0)𝑛𝑥0n(x=0)italic_n ( italic_x = 0 ) at the center of the barrier are used. From the good agreement between these two curves we are led to conclude that it is the dissipative mechanism of the back-flow current which develops inside the barrier to determine the value of the critical current for the Josephson junction. Note how, in both cases, the maximum (critical) currents correspond to the Bardeen critical current discussed in Sec. III.1.

IV Decay of persistent currents

A topic for which the concept of intrinsic critical current discussed above is especially relevant is that of the persistence of a super-current induced in a superfluid with a closed (annular) geometry and of the associated decay mechanisms, which are the hallmark of superfluidity in the first place. In principle, a persistent current will continue indefinitely as long as the medium is superfluid. In practice, persistent currents in superconducting materials (like, for instance, NbZr alloys) were estimated to flow over more than hundreds of years File-Mills-1963 . Recently, this topic was taken over in the context of ultra-cold Fermi gases, for which a recent experiment DelPace-2022 using a phase-imprinting technique has detected quantized circulations across the BCS-BEC crossover persisting up to a few seconds and identified its decay mechanism to take place via the emission of vortices Xhani-2023 . In this Section, we examine the outcomes of this experiment in the light of the theoretical considerations made above in Sec. III.

Refer to caption
Figure 7: Experimental data for the final mean winding number wFdelimited-⟨⟩subscript𝑤𝐹\langle w_{F}\rangle⟨ italic_w start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ⟩ vs the impressed winding number w0delimited-⟨⟩subscript𝑤0\langle w_{0}\rangle⟨ italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ are compared with theoretical calculations for couplings (kFaF)1=0.4superscriptsubscript𝑘𝐹subscript𝑎𝐹10.4(k_{F}a_{F})^{-1}=-0.4( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = - 0.4 (top panel) and (kFaF)1=0.0superscriptsubscript𝑘𝐹subscript𝑎𝐹10.0(k_{F}a_{F})^{-1}=0.0( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = 0.0 (bottom panel). Experimental data for a homogeneous superfluid (diamonds) are from Fig. 3(c) and those in presence of a point defect (triangles) from Fig. 4(e) of Ref. DelPace-2022 . The results of the theoretical simulations at temperature T=0.02TF𝑇0.02subscript𝑇𝐹T=0.02T_{F}italic_T = 0.02 italic_T start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT (as described in detail in the text) are identified by filled dots up to the critical value of wFdelimited-⟨⟩subscript𝑤𝐹\langle w_{F}\rangle⟨ italic_w start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ⟩ and by empty dots past this critical value.

Typically, along a circle of radius R𝑅Ritalic_R the velocity field 𝐯=φ(𝐫)/(2m)𝐯𝜑𝐫2𝑚\mathbf{v}=\nabla\varphi(\mathbf{r})/(2m)bold_v = ∇ italic_φ ( bold_r ) / ( 2 italic_m ) of the fermionic superfluid (where φ(𝐫)𝜑𝐫\varphi(\mathbf{r})italic_φ ( bold_r ) is the phase of the local gap parameter Δ(𝐫)=|Δ(𝐫)|eiφ(𝐫)Δ𝐫Δ𝐫superscript𝑒𝑖𝜑𝐫\Delta(\mathbf{r})=|\Delta(\mathbf{r})|e^{i\varphi(\mathbf{r})}roman_Δ ( bold_r ) = | roman_Δ ( bold_r ) | italic_e start_POSTSUPERSCRIPT italic_i italic_φ ( bold_r ) end_POSTSUPERSCRIPT and m𝑚mitalic_m is the fermion mass) has a quantized circulation given by πw/m𝜋𝑤𝑚\pi w/mitalic_π italic_w / italic_m where w𝑤witalic_w is an integer. Correspondingly, the magnitude of 𝐯𝐯\mathbf{v}bold_v takes the quantized values v=w/(2mR)𝑣𝑤2𝑚𝑅v=w/(2mR)italic_v = italic_w / ( 2 italic_m italic_R ). The integer w𝑤witalic_w is referred to as a winding number just because it counts the number of oscillations of the phase of the order parameter along the circulation. It can be measured by interferometric techniques. After an initial overall phase difference 2πw02𝜋subscript𝑤02\pi w_{0}2 italic_π italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT has been imprinted on the superfluid, the system is observed to stabilize for a time of the order of tenths of a second and its final (mean) winding number wFdelimited-⟨⟩subscript𝑤𝐹\langle w_{F}\rangle⟨ italic_w start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ⟩ is measured. Experimentally, mean winding numbers (both for w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and wFsubscript𝑤𝐹w_{F}italic_w start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT) are considered because the same imprinting procedure is repeated several times, such that the ensuing results are suitably averaged out.

Figure 7 reports the experimental data from Ref. DelPace-2022 for the mean winding number wFdelimited-⟨⟩subscript𝑤𝐹\langle w_{F}\rangle⟨ italic_w start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ⟩ at low temperature for the couplings (kFaF)1=(0.4,0.0)superscriptsubscript𝑘𝐹subscript𝑎𝐹10.40.0(k_{F}a_{F})^{-1}=(-0.4,0.0)( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = ( - 0.4 , 0.0 ), as obtained for a homogeneous superfluid constrained in a ring geometry (diamonds) and with the additional presence of a weak point-like defect placed inside the ring (triangles). In the experiment dissipation is expected to occur as soon as wF<w0delimited-⟨⟩subscript𝑤𝐹delimited-⟨⟩subscript𝑤0\langle w_{F}\rangle<\langle w_{0}\rangle⟨ italic_w start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ⟩ < ⟨ italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩. However, while the onset of dissipation is observed in both geometries for the weaker coupling (kFaF)1=0.4superscriptsubscript𝑘𝐹subscript𝑎𝐹10.4(k_{F}a_{F})^{-1}=-0.4( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = - 0.4 as shown in Fig. 7(a), at unitarity apparently no dissipation is found in either geometry as shown in Fig. 7(b). Here, we provide a theoretical explanation for this experimental finding in the following terms.

We first unfold the ring into a linear tube of the same length and supplement it with periodic boundary conditions at its edges. We then adopt a strategy similar to that recently utilized in Ref. Piselli-2023 in terms of the mLPDA approach (see Appendix A below) and partition the tube into a large number of (961) tubular filaments, each of which is treated as if it were a homogeneous superfluid with a given (local) density and a linear super-current flowing through it. To have full control of the density profiles spanning these filaments, we adjust the number of atoms as well as the height and width of the walls that contain the atomic cloud, in such a way to reproduce the corresponding experimental density profiles. A comparison between the experimental and theoretical density profiles obtained in this way is shown in Fig. 8 for the same couplings considered in Fig. 7.

With this calibration procedure at hand, we may now restrict ourselves to considering the tubular filament that corresponds to the innermost part of the original ring with the smallest distance Rinsubscript𝑅inR_{\mathrm{in}}italic_R start_POSTSUBSCRIPT roman_in end_POSTSUBSCRIPT from its center where the critical flow is reached first. This is a consequence of the facts that all filaments must have the same winding number and that the quantization for the velocity obtained above reads q=w/(2R)𝑞𝑤2𝑅q=w/(2R)italic_q = italic_w / ( 2 italic_R ) in terms of the linear momentum q=v/m𝑞𝑣𝑚q=v/mitalic_q = italic_v / italic_m at a distance R𝑅Ritalic_R from the ring center. In this way, the intrinsic current density j(q0)𝑗subscript𝑞0j(q_{0})italic_j ( italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) is computed for several momenta q0=w0/(2Rin)subscript𝑞0subscript𝑤02subscript𝑅inq_{0}=w_{0}/(2R_{\mathrm{in}})italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / ( 2 italic_R start_POSTSUBSCRIPT roman_in end_POSTSUBSCRIPT ) along the lines of Sec. III.1, and the final winding number wFsubscript𝑤𝐹w_{F}italic_w start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT is obtained from the relations qF/m=j(q0)/n(Rin)subscript𝑞𝐹𝑚𝑗subscript𝑞0𝑛subscript𝑅inq_{F}/m=j(q_{0})/n(R_{\mathrm{in}})italic_q start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT / italic_m = italic_j ( italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / italic_n ( italic_R start_POSTSUBSCRIPT roman_in end_POSTSUBSCRIPT ) and qF=wF/(2Rin)subscript𝑞𝐹subscript𝑤𝐹2subscript𝑅inq_{F}=w_{F}/(2R_{\mathrm{in}})italic_q start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = italic_w start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT / ( 2 italic_R start_POSTSUBSCRIPT roman_in end_POSTSUBSCRIPT ) where n(Rin)𝑛subscript𝑅inn(R_{\mathrm{in}})italic_n ( italic_R start_POSTSUBSCRIPT roman_in end_POSTSUBSCRIPT ) is the local particle density at position Rinsubscript𝑅inR_{\mathrm{in}}italic_R start_POSTSUBSCRIPT roman_in end_POSTSUBSCRIPT. The results of this calculation are then reported in Fig. 7, where they are compared with the experimental data of Ref. DelPace-2022 .

Refer to caption
Figure 8: Comparison between experimental (full lines) and theoretical (dashed lines) density profiles with the ring geometry of Ref. DelPace-2022 for couplings (kFaF)1=0.4superscriptsubscript𝑘𝐹subscript𝑎𝐹10.4(k_{F}a_{F})^{-1}=-0.4( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = - 0.4 (top panel) and (kFaF)1=0.0superscriptsubscript𝑘𝐹subscript𝑎𝐹10.0(k_{F}a_{F})^{-1}=0.0( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = 0.0 (bottom panel). [Experimental profiles courtesy of G. Del Pace.]

Note from this figure that, as long as the momentum q0subscript𝑞0q_{0}italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is below the critical momentum corresponding to the Landau dissipative threshold, the initial w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and final wFsubscript𝑤𝐹w_{F}italic_w start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT winding numbers coincide with each other as expected. Conversely, above this critical momentum (identified by the end point of the linear increase in Fig. 7) the onset of the back-flow current acts to suppress the value of the super-current, thus making wF<w0subscript𝑤𝐹subscript𝑤0w_{F}<w_{0}italic_w start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT < italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Note also that the presence of a weak barrier on the experimental determination of the winding number has only a minor effect for coupling (kFaF)1=0.4superscriptsubscript𝑘𝐹subscript𝑎𝐹10.4(k_{F}a_{F})^{-1}=-0.4( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = - 0.4 and essentially no effect for coupling (kFaF)1=0.0superscriptsubscript𝑘𝐹subscript𝑎𝐹10.0(k_{F}a_{F})^{-1}=0.0( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = 0.0. To confirm this experimental finding, we have also explicitly considered the presence of a weak barrier in our theoretical calculations, although only to realize that its effects are indeed quite negligible. Note, finally, that Fig. 7 predicts that, at unitarity, the theoretical value of the threshold at which wFsubscript𝑤𝐹w_{F}italic_w start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT deviates from a linear behavior considerably exceeds the maximum value of w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT experimentally attainable.

V Conclusions

In this article, we have performed a systematic investigation of the current-vs-momentum characteristics for a fermionic superfluid spanning the BCS-BEC crossover, for which the inclusion of pairing fluctuations beyond mean field plays a crucial role. To this end, we have initially considered a fully homogeneous superfluid system, for which the above characteristics have allowed us to identify the intrinsic upper value of the super-current flowing through the system, in terms of a single expression of the super-current which is intrinsically limited by two different dissipative mechanisms (pair-breaking and phonon excitations) on the two BCS and BEC sides of the crossover, respectively. In this context, we have considered both the Landau Landau-1941 and Bardeen Bardeen-1962 criteria for the loss of superfluid behavior when the intrinsic current exceeds a corresponding upper value, and examined how their respective values evolve across the BCS-BEC crossover within suitable temperature ranges. When needed, we have further included the effects of a small barrier that somewhat spoils the system homogeneity, so as to reproduce as closely as possible the experimental configurations based on linear Weimer-2015 and annular DelPace-2022 geometries.

In our analysis, we have not explicitly investigated the fate of the homogeneous superfluid once passed the Landau or Bardeen thresholds, when one expects patterns of inhomogeneous fluctuations to be built from pair-breaking (on the BCS side) and sound (on the BEC side) elementary excitations, giving rise to spatially non-uniform configurations with a strong local suppression of the gap parameter. Typically, these configurations are expected to be precursors for the formation of vortices. This topic, although interesting in itself both theoretically and experimentally, exceeds the purposes of the present article. In this respect, one may envisage that an appropriate theoretical method should possibly be based on a time-dependent approach, that would include pairing fluctuations and spatially inhomogeneous configurations on the same footing over a wide range of temperature and coupling throughout the BCS-BEC crossover.

Acknowledgments - We are indebted to G. Roati and G. Del Pace for a critical reading of the manuscript.

Appendix A BRIEF OVERVIEW OF THE MAIN
PREVIOUS THEORETICAL RESULTS
UTILIZED IN THE PRESENT WORK

For the benefit of the readers, this Appendix briefly summarizes a number of theoretical results previously obtained in the literature, which are relevant to obtain the results discussed in the present work for the critical current with the inclusion of pairing fluctuations beyond mean field that span the BCS-BEC crossover. The results here summarized include (i) the non-self-consistent t𝑡titalic_t-matrix approach in the presence of a superfluid flow that was implemented in Ref. Pisani-2023 , (ii) the modified Local Phase Density Approximation (mLPDA) that was also introduced in Ref. Pisani-2023 to include pairing fluctuations over and above the original Local Phase Density Approximation (LPDA) of Ref. Simonucci-2014 , and (iii) the extended Gorkov-Melik-Barkhudarov (GMB) approach originally introduced in Refs. Pisani-2018-a and Pisani-2018-b .

1. Expressions for the density and current

in the presence of a superfluid flow

In the superfluid phase, the fermionic local number density and current read

n(𝐫)=2βneiωnηG11(𝐫,𝐫;ωn)𝑛𝐫2𝛽subscript𝑛superscript𝑒𝑖subscript𝜔𝑛𝜂subscript𝐺11𝐫𝐫subscript𝜔𝑛n(\mathbf{r})=\dfrac{2}{\beta}\sum_{n}e^{i\omega_{n}\eta}G_{11}(\mathbf{r},% \mathbf{r};\omega_{n})italic_n ( bold_r ) = divide start_ARG 2 end_ARG start_ARG italic_β end_ARG ∑ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_η end_POSTSUPERSCRIPT italic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( bold_r , bold_r ; italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) (14a)
j(𝐫)=1βneiωnη(𝐫𝐫)imG11(𝐫,𝐫;ωn)|𝐫=𝐫,j(\mathbf{r})=\dfrac{1}{\beta}\sum_{n}e^{i\omega_{n}\eta}\dfrac{\left(\nabla_{% \mathbf{r}}-\nabla_{\mathbf{r^{\prime}}}\right)}{im}G_{11}(\mathbf{r},\mathbf{% r^{\prime}};\omega_{n})\rvert_{\mathbf{r}=\mathbf{r^{\prime}}}\,,italic_j ( bold_r ) = divide start_ARG 1 end_ARG start_ARG italic_β end_ARG ∑ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_η end_POSTSUPERSCRIPT divide start_ARG ( ∇ start_POSTSUBSCRIPT bold_r end_POSTSUBSCRIPT - ∇ start_POSTSUBSCRIPT bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ) end_ARG start_ARG italic_i italic_m end_ARG italic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( bold_r , bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ; italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) | start_POSTSUBSCRIPT bold_r = bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT , (14b)

where G11subscript𝐺11G_{11}italic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT is the “normal” single-particle Green’s function, β=(kBT)1𝛽superscriptsubscript𝑘𝐵𝑇1\beta=(k_{B}T)^{-1}italic_β = ( italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT the inverse temperature (kBsubscript𝑘𝐵k_{B}italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT being the Boltzmann constant), η𝜂\etaitalic_η a positive infinitesimal, m𝑚mitalic_m the fermion mass, and ωn=(2n+1)π/βsubscript𝜔𝑛2𝑛1𝜋𝛽\omega_{n}=(2n+1)\pi/\betaitalic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = ( 2 italic_n + 1 ) italic_π / italic_β (n𝑛nitalic_n integer) a fermionic Matsubara frequency Schrieffer-1964 .

When a super-current with momentum 𝐪𝐪\mathbf{q}bold_q flows in a homogeneous environment, the gap parameter takes the form DeGennes-1966

Δ(𝐫)=ei2𝐪𝐫Δ𝐪.Δ𝐫superscript𝑒𝑖2𝐪𝐫subscriptΔ𝐪\Delta(\mathbf{r})=e^{i2\mathbf{q}\cdot\mathbf{r}}\Delta_{\mathbf{q}}\,.roman_Δ ( bold_r ) = italic_e start_POSTSUPERSCRIPT italic_i 2 bold_q ⋅ bold_r end_POSTSUPERSCRIPT roman_Δ start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT . (15)

To comply with this spatial dependence, the single-particle Green’s function G11subscript𝐺11G_{11}italic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT then becomes Pisani-2023

G11(x,x;𝐪)=ei𝐪(𝐫𝐫)𝒢11(xx;𝐪)subscript𝐺11𝑥superscript𝑥𝐪superscript𝑒𝑖𝐪𝐫superscript𝐫subscript𝒢11𝑥superscript𝑥𝐪G_{11}(x,x^{\prime};\mathbf{q})=e^{i\mathbf{q}\cdot(\mathbf{r}-\mathbf{r^{% \prime}})}\mathcal{G}_{11}(x-x^{\prime};\mathbf{q})italic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( italic_x , italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ; bold_q ) = italic_e start_POSTSUPERSCRIPT italic_i bold_q ⋅ ( bold_r - bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) end_POSTSUPERSCRIPT caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( italic_x - italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ; bold_q ) (16)

where 𝒢11subscript𝒢11\mathcal{G}_{11}caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT is the “reduced” single-particle Green’s function

𝒢11(xx;𝐪)=kei𝐤(𝐫𝐫)eiωn(ττ)𝒢11(k;𝐪)subscript𝒢11𝑥superscript𝑥𝐪subscript𝑘superscript𝑒𝑖𝐤𝐫superscript𝐫superscript𝑒𝑖subscript𝜔𝑛𝜏superscript𝜏subscript𝒢11𝑘𝐪\mathcal{G}_{11}(x-x^{\prime};\mathbf{q})=\sum_{k}e^{i\mathbf{k}\cdot(\mathbf{% r}-\mathbf{r^{\prime}})}e^{-i\omega_{n}(\tau-\tau^{\prime})}\mathcal{G}_{11}(k% ;\mathbf{q})caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( italic_x - italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ; bold_q ) = ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i bold_k ⋅ ( bold_r - bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_τ - italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) end_POSTSUPERSCRIPT caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( italic_k ; bold_q ) (17)

with the short-hand notation k=d𝐤(2π)31βnsubscript𝑘𝑑𝐤superscript2𝜋31𝛽subscript𝑛\sum_{k}=\int\dfrac{d\mathbf{k}}{(2\pi)^{3}}\dfrac{1}{\beta}\sum_{n}∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = ∫ divide start_ARG italic_d bold_k end_ARG start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG divide start_ARG 1 end_ARG start_ARG italic_β end_ARG ∑ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, k=(𝐤,ωn)𝑘𝐤subscript𝜔𝑛k=(\mathbf{k},\omega_{n})italic_k = ( bold_k , italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) being a fermionic four-vector. Accordingly, in momentum and frequency space the expressions (14a) and (14b) read

n=2d𝐤(2π)31βneiωnη𝒢11(𝐤,ωn;𝐪)𝑛2𝑑𝐤superscript2𝜋31𝛽subscript𝑛superscript𝑒𝑖subscript𝜔𝑛𝜂subscript𝒢11𝐤subscript𝜔𝑛𝐪n=2\int\frac{d\mathbf{k}}{(2\pi)^{3}}\frac{1}{\beta}\sum_{n}e^{i\omega_{n}\eta% }\mathcal{G}_{11}\left(\mathbf{k},\omega_{n};\mathbf{q}\right)italic_n = 2 ∫ divide start_ARG italic_d bold_k end_ARG start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG divide start_ARG 1 end_ARG start_ARG italic_β end_ARG ∑ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_η end_POSTSUPERSCRIPT caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( bold_k , italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ; bold_q ) (18a)
𝐣=𝐪mn+2d𝐤(2π)3𝐤m1βneiωnη𝒢11(𝐤,ωn;𝐪).𝐣𝐪𝑚𝑛2𝑑𝐤superscript2𝜋3𝐤𝑚1𝛽subscript𝑛superscript𝑒𝑖subscript𝜔𝑛𝜂subscript𝒢11𝐤subscript𝜔𝑛𝐪\mathbf{j}=\frac{\mathbf{q}}{m}n+2\int\frac{d\mathbf{k}}{(2\pi)^{3}}\frac{% \mathbf{k}}{m}\frac{1}{\beta}\sum_{n}e^{i\omega_{n}\eta}\mathcal{G}_{11}\left(% \mathbf{k},\omega_{n};\mathbf{q}\right)\,.bold_j = divide start_ARG bold_q end_ARG start_ARG italic_m end_ARG italic_n + 2 ∫ divide start_ARG italic_d bold_k end_ARG start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG divide start_ARG bold_k end_ARG start_ARG italic_m end_ARG divide start_ARG 1 end_ARG start_ARG italic_β end_ARG ∑ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_η end_POSTSUPERSCRIPT caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( bold_k , italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ; bold_q ) . (18b)

The form of 𝒢11(k;𝐪)subscript𝒢11𝑘𝐪\mathcal{G}_{11}\left(k;\mathbf{q}\right)caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( italic_k ; bold_q ) to be entered in the above expressions can be evaluated within different approximations, depending on the choice of the single-particle self-energy. In the present work, it is evaluated both at the mean-field level and with the inclusion of pairing fluctuations within the non-self-consistent t𝑡titalic_t-matrix approach. Quite generally, 𝒢11subscript𝒢11\mathcal{G}_{11}caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT is obtained by solving the Dyson’s equation in the broken-symmetry phase in the Nambu-Gorkov formalism Pisani-2023

[iωnξ(𝐤+𝐪)𝔖11(k;𝐪)𝔖12(k;𝐪)𝔖21(k;𝐪)iωn+ξ(𝐤𝐪)𝔖22(k;𝐪)][𝒢11(k;𝐪)𝒢12(k;𝐪)𝒢21(k;𝐪)𝒢22(k;𝐪)]=[1001],matrix𝑖subscript𝜔𝑛𝜉𝐤𝐪subscript𝔖11𝑘𝐪subscript𝔖12𝑘𝐪subscript𝔖21𝑘𝐪𝑖subscript𝜔𝑛𝜉𝐤𝐪subscript𝔖22𝑘𝐪matrixsubscript𝒢11𝑘𝐪subscript𝒢12𝑘𝐪subscript𝒢21𝑘𝐪subscript𝒢22𝑘𝐪matrix1001\begin{bmatrix}i\omega_{n}-\xi(\mathbf{k}+\mathbf{q})-\mathfrak{S}_{11}(k;% \mathbf{q})&-\mathfrak{S}_{12}(k;\mathbf{q})\\ -\mathfrak{S}_{21}(k;\mathbf{q})&i\omega_{n}+\xi(\mathbf{k}-\mathbf{q})-% \mathfrak{S}_{22}(k;\mathbf{q})\end{bmatrix}\begin{bmatrix}\mathcal{G}_{11}(k;% \mathbf{q})&\mathcal{G}_{12}(k;\mathbf{q})\\ \mathcal{G}_{21}(k;\mathbf{q})&\mathcal{G}_{22}(k;\mathbf{q})\end{bmatrix}=% \begin{bmatrix}1&0\\ 0&1\end{bmatrix}\,,[ start_ARG start_ROW start_CELL italic_i italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - italic_ξ ( bold_k + bold_q ) - fraktur_S start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( italic_k ; bold_q ) end_CELL start_CELL - fraktur_S start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT ( italic_k ; bold_q ) end_CELL end_ROW start_ROW start_CELL - fraktur_S start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT ( italic_k ; bold_q ) end_CELL start_CELL italic_i italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + italic_ξ ( bold_k - bold_q ) - fraktur_S start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT ( italic_k ; bold_q ) end_CELL end_ROW end_ARG ] [ start_ARG start_ROW start_CELL caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( italic_k ; bold_q ) end_CELL start_CELL caligraphic_G start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT ( italic_k ; bold_q ) end_CELL end_ROW start_ROW start_CELL caligraphic_G start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT ( italic_k ; bold_q ) end_CELL start_CELL caligraphic_G start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT ( italic_k ; bold_q ) end_CELL end_ROW end_ARG ] = [ start_ARG start_ROW start_CELL 1 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL 1 end_CELL end_ROW end_ARG ] , (19)

where ξ(𝐤)=𝐤2/(2m)μ𝜉𝐤superscript𝐤22𝑚𝜇\xi(\mathbf{k})=\mathbf{k}^{2}/(2m)-\muitalic_ξ ( bold_k ) = bold_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 2 italic_m ) - italic_μ and 𝔖iisubscript𝔖𝑖superscript𝑖\mathfrak{S}_{ii^{\prime}}fraktur_S start_POSTSUBSCRIPT italic_i italic_i start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT are the components of the “reduced” self-energy. Different approximations then correspond to different choices of 𝔖iisubscript𝔖𝑖superscript𝑖\mathfrak{S}_{ii^{\prime}}fraktur_S start_POSTSUBSCRIPT italic_i italic_i start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT.

1-a - Mean-field approximation

in the presence of a super-current

At the mean-field (mfmf\mathrm{mf}roman_mf) level, one takes 𝔖11(k,𝐪)=𝔖22(k,𝐪)=0subscript𝔖11𝑘𝐪subscript𝔖22𝑘𝐪0\mathfrak{S}_{11}(k,\mathbf{q})=\mathfrak{S}_{22}(k,\mathbf{q})=0fraktur_S start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( italic_k , bold_q ) = fraktur_S start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT ( italic_k , bold_q ) = 0 and 𝔖12(k,𝐪)=𝔖21(k,𝐪)=Δ𝐪subscript𝔖12𝑘𝐪subscript𝔖21𝑘𝐪subscriptΔ𝐪\mathfrak{S}_{12}(k,\mathbf{q})=\mathfrak{S}_{21}(k,\mathbf{q})=-\Delta_{% \mathbf{q}}fraktur_S start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT ( italic_k , bold_q ) = fraktur_S start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT ( italic_k , bold_q ) = - roman_Δ start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT, such that solving the Dyson’s equation (19) yields Pisani-2023

𝒢11mf(k;𝐪)=u(𝐤;𝐪)2iωnE+(𝐤;𝐪)+v(𝐤;𝐪)2iωn+E(𝐤;𝐪)superscriptsubscript𝒢11mf𝑘𝐪𝑢superscript𝐤𝐪2𝑖subscript𝜔𝑛subscript𝐸𝐤𝐪𝑣superscript𝐤𝐪2𝑖subscript𝜔𝑛subscript𝐸𝐤𝐪\mathcal{G}_{11}^{\mathrm{mf}}(k;\mathbf{q})=\dfrac{u(\mathbf{k};\mathbf{q})^{% 2}}{i\omega_{n}-E_{+}(\mathbf{k};\mathbf{q})}+\dfrac{v(\mathbf{k};\mathbf{q})^% {2}}{i\omega_{n}+E_{-}(\mathbf{k};\mathbf{q})}caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_mf end_POSTSUPERSCRIPT ( italic_k ; bold_q ) = divide start_ARG italic_u ( bold_k ; bold_q ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_i italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ( bold_k ; bold_q ) end_ARG + divide start_ARG italic_v ( bold_k ; bold_q ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_i italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + italic_E start_POSTSUBSCRIPT - end_POSTSUBSCRIPT ( bold_k ; bold_q ) end_ARG (20a)
𝒢12mf(k;𝐪)=u(k;𝐪)v(k;𝐪)×[1iωnE+(𝐤;𝐪)1iωn+E(𝐤;𝐪)],superscriptsubscript𝒢12mf𝑘𝐪𝑢𝑘𝐪𝑣𝑘𝐪delimited-[]1𝑖subscript𝜔𝑛subscript𝐸𝐤𝐪1𝑖subscript𝜔𝑛subscript𝐸𝐤𝐪\begin{split}\mathcal{G}_{12}^{\mathrm{mf}}(k;\mathbf{q})&=-u(k;\mathbf{q})v(k% ;\mathbf{q})\\ &\times\left[\dfrac{1}{i\omega_{n}-E_{+}(\mathbf{k};\mathbf{q})}-\dfrac{1}{i% \omega_{n}+E_{-}(\mathbf{k};\mathbf{q})}\right],\end{split}start_ROW start_CELL caligraphic_G start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_mf end_POSTSUPERSCRIPT ( italic_k ; bold_q ) end_CELL start_CELL = - italic_u ( italic_k ; bold_q ) italic_v ( italic_k ; bold_q ) end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL × [ divide start_ARG 1 end_ARG start_ARG italic_i italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ( bold_k ; bold_q ) end_ARG - divide start_ARG 1 end_ARG start_ARG italic_i italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + italic_E start_POSTSUBSCRIPT - end_POSTSUBSCRIPT ( bold_k ; bold_q ) end_ARG ] , end_CELL end_ROW (20b)

where

u(𝐤;𝐪)2=12(1+ξ(𝐤;𝐪)E(𝐤;𝐪))𝑢superscript𝐤𝐪2121𝜉𝐤𝐪𝐸𝐤𝐪u(\mathbf{k};\mathbf{q})^{2}=\dfrac{1}{2}\left(1+\dfrac{\xi(\mathbf{k};\mathbf% {q})}{E(\mathbf{k};\mathbf{q})}\right)italic_u ( bold_k ; bold_q ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( 1 + divide start_ARG italic_ξ ( bold_k ; bold_q ) end_ARG start_ARG italic_E ( bold_k ; bold_q ) end_ARG ) (21a)
v(𝐤;𝐪)2=12(1ξ(𝐤;𝐪)E(𝐤;𝐪))𝑣superscript𝐤𝐪2121𝜉𝐤𝐪𝐸𝐤𝐪v(\mathbf{k};\mathbf{q})^{2}=\dfrac{1}{2}\left(1-\dfrac{\xi(\mathbf{k};\mathbf% {q})}{E(\mathbf{k};\mathbf{q})}\right)italic_v ( bold_k ; bold_q ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( 1 - divide start_ARG italic_ξ ( bold_k ; bold_q ) end_ARG start_ARG italic_E ( bold_k ; bold_q ) end_ARG ) (21b)

with the notation

ξ(𝐤;𝐪)=𝐤22mμ+𝐪22mE(𝐤;𝐪)=ξ(𝐤;𝐪)2+Δ𝐪2E±(𝐤;𝐪)=E(𝐤;𝐪)±𝐤𝐪m.𝜉𝐤𝐪superscript𝐤22𝑚𝜇superscript𝐪22𝑚𝐸𝐤𝐪𝜉superscript𝐤𝐪2superscriptsubscriptΔ𝐪2subscript𝐸plus-or-minus𝐤𝐪plus-or-minus𝐸𝐤𝐪𝐤𝐪𝑚\begin{split}\xi(\mathbf{k};\mathbf{q})&=\dfrac{\mathbf{k}^{2}}{2m}-\mu+\dfrac% {\mathbf{q}^{2}}{2m}\\ E(\mathbf{k};\mathbf{q})&=\sqrt{\xi(\mathbf{k};\mathbf{q})^{2}+\Delta_{\mathbf% {q}}^{2}}\\ E_{\pm}(\mathbf{k};\mathbf{q})&=E(\mathbf{k};\mathbf{q})\pm\dfrac{\mathbf{k}% \cdot\mathbf{q}}{m}\,.\end{split}start_ROW start_CELL italic_ξ ( bold_k ; bold_q ) end_CELL start_CELL = divide start_ARG bold_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG - italic_μ + divide start_ARG bold_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG end_CELL end_ROW start_ROW start_CELL italic_E ( bold_k ; bold_q ) end_CELL start_CELL = square-root start_ARG italic_ξ ( bold_k ; bold_q ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Δ start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_CELL end_ROW start_ROW start_CELL italic_E start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ( bold_k ; bold_q ) end_CELL start_CELL = italic_E ( bold_k ; bold_q ) ± divide start_ARG bold_k ⋅ bold_q end_ARG start_ARG italic_m end_ARG . end_CELL end_ROW (22)

Entering the result (20a) for 𝒢11subscript𝒢11\mathcal{G}_{11}caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT into the expressions (A), one ends up with the expression (1) for the current density at the mean-filed level (as well as with the associated expression for the density), which has the typical form of the two-fluid model at finite temperature in the Bardeen formulation for fermions Bardeen-1962 .

1-b - t𝑡titalic_t-matrix approximation

in the presence of a super-current

Within the pairing-fluctuation (pfpf\mathrm{pf}roman_pf) approximation, the reduced self-energy of the Dyson’s equation (19) reads Pisani-2023

{𝔖11pf(k;𝐪)=𝔖22pf(k;𝐪)=QΓ11(Q;𝐪)𝒢11mf(Qk;𝐪)𝔖12pf(k;q)=𝔖21pf(k;q)=Δ𝐪casessuperscriptsubscript𝔖11pf𝑘𝐪superscriptsubscript𝔖22pf𝑘𝐪absentsubscript𝑄subscriptΓ11𝑄𝐪superscriptsubscript𝒢11mf𝑄𝑘𝐪missing-subexpressionsuperscriptsubscript𝔖12pf𝑘𝑞superscriptsubscript𝔖21pf𝑘𝑞subscriptΔ𝐪\displaystyle\left\{\begin{array}[]{l}\mathfrak{S}_{11}^{\mathrm{pf}}(k;{\bf q% })=-\mathfrak{S}_{22}^{\mathrm{pf}}(-k;{\bf q})\\ \hskip 44.10185pt=-\sum_{Q}\Gamma_{11}(Q;{\bf q})\mathcal{G}_{11}^{\mathrm{mf}% }(Q-k;{\bf q})\\ \\ \mathfrak{S}_{12}^{\mathrm{pf}}(k;q)=\mathfrak{S}_{21}^{\mathrm{pf}}(k;q)=-% \Delta_{\bf q}\end{array}\right.{ start_ARRAY start_ROW start_CELL fraktur_S start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_pf end_POSTSUPERSCRIPT ( italic_k ; bold_q ) = - fraktur_S start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_pf end_POSTSUPERSCRIPT ( - italic_k ; bold_q ) end_CELL end_ROW start_ROW start_CELL = - ∑ start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT roman_Γ start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( italic_Q ; bold_q ) caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_mf end_POSTSUPERSCRIPT ( italic_Q - italic_k ; bold_q ) end_CELL end_ROW start_ROW start_CELL end_CELL end_ROW start_ROW start_CELL fraktur_S start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_pf end_POSTSUPERSCRIPT ( italic_k ; italic_q ) = fraktur_S start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_pf end_POSTSUPERSCRIPT ( italic_k ; italic_q ) = - roman_Δ start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT end_CELL end_ROW end_ARRAY (27)

with the short-hand notation Qd𝐐(2π)31βνsubscript𝑄𝑑𝐐superscript2𝜋31𝛽subscript𝜈\sum_{Q}\longleftrightarrow\int\frac{d{\bf Q}}{(2\pi)^{3}}\frac{1}{\beta}\sum_% {\nu}∑ start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ⟷ ∫ divide start_ARG italic_d bold_Q end_ARG start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG divide start_ARG 1 end_ARG start_ARG italic_β end_ARG ∑ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT, where Q=(𝐐,Ωv)𝑄𝐐subscriptΩ𝑣Q=\left({\bf Q},\Omega_{v}\right)italic_Q = ( bold_Q , roman_Ω start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ) is a four-vector with Ων=2νπ/βsubscriptΩ𝜈2𝜈𝜋𝛽\Omega_{\nu}=2\nu\pi/\betaroman_Ω start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT = 2 italic_ν italic_π / italic_β (ν𝜈\nuitalic_ν integer) a bosonic Matsubara frequency FW-1971 . In this way, the expression of 𝒢11(k;𝐪)subscript𝒢11𝑘𝐪\mathcal{G}_{11}\left(k;\mathbf{q}\right)caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( italic_k ; bold_q ), to be utilized in Eq. (18b) to obtain the current with the inclusion of pairing fluctuations, becomes:

𝒢11pf(k;𝐪)=1iωnξ(𝐤+𝐪)𝔖11pf(k;𝐪)Δ𝐪2iωn+ξ(𝐤𝐪)+𝔖11pf(k;𝐪).superscriptsubscript𝒢11pf𝑘𝐪1𝑖subscript𝜔𝑛𝜉𝐤𝐪superscriptsubscript𝔖11pf𝑘𝐪superscriptsubscriptΔ𝐪2𝑖subscript𝜔𝑛𝜉𝐤𝐪superscriptsubscript𝔖11pf𝑘𝐪\mathcal{G}_{11}^{\mathrm{pf}}(k;\mathbf{q})=\frac{1}{i\omega_{n}-\xi(\mathbf{% k}+\mathbf{q})-\mathfrak{S}_{11}^{\mathrm{pf}}(k;\mathbf{q})-\frac{\Delta_{% \mathbf{q}}^{2}}{i\omega_{n}+\xi(\mathbf{k}-\mathbf{q})+\mathfrak{S}_{11}^{% \mathrm{pf}}(-k;\mathbf{q})}}\,.caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_pf end_POSTSUPERSCRIPT ( italic_k ; bold_q ) = divide start_ARG 1 end_ARG start_ARG italic_i italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - italic_ξ ( bold_k + bold_q ) - fraktur_S start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_pf end_POSTSUPERSCRIPT ( italic_k ; bold_q ) - divide start_ARG roman_Δ start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_i italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + italic_ξ ( bold_k - bold_q ) + fraktur_S start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_pf end_POSTSUPERSCRIPT ( - italic_k ; bold_q ) end_ARG end_ARG . (28)

The quantity Γ11subscriptΓ11\Gamma_{11}roman_Γ start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT entering Eq. (27) is the upper diagonal element of the 2×2222\times 22 × 2 matrix for the “pair propagator”, which consists in a series of ladder diagrams whereby two fermions with opposite spins repeatedly scatter with each other Andrenacci-2003 . In the presence of a super-current, its expression reads Pisani-2023

[Γ11(Q;𝐪)Γ12(Q;𝐪)Γ21(Q;𝐪)Γ22(Q;𝐪)]delimited-[]subscriptΓ11𝑄𝐪subscriptΓ12𝑄𝐪subscriptΓ21𝑄𝐪subscriptΓ22𝑄𝐪\displaystyle{\left[\begin{array}[]{ll}\Gamma_{11}(Q;{\bf q})&\Gamma_{12}(Q;{% \bf q})\\ \Gamma_{21}(Q;{\bf q})&\Gamma_{22}(Q;{\bf q})\end{array}\right]}[ start_ARRAY start_ROW start_CELL roman_Γ start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( italic_Q ; bold_q ) end_CELL start_CELL roman_Γ start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT ( italic_Q ; bold_q ) end_CELL end_ROW start_ROW start_CELL roman_Γ start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT ( italic_Q ; bold_q ) end_CELL start_CELL roman_Γ start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT ( italic_Q ; bold_q ) end_CELL end_ROW end_ARRAY ] =1A(Q;𝐪)A(Q;𝐪)B(Q;𝐪)2absent1𝐴𝑄𝐪𝐴𝑄𝐪𝐵superscript𝑄𝐪2\displaystyle=\frac{1}{A(Q;{\bf q})A(-Q;{\bf q})-B(Q;{\bf q})^{2}}= divide start_ARG 1 end_ARG start_ARG italic_A ( italic_Q ; bold_q ) italic_A ( - italic_Q ; bold_q ) - italic_B ( italic_Q ; bold_q ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG (31)
×[A(Q;𝐪)B(Q;𝐪)B(Q;𝐪)A(Q;𝐪)]absentdelimited-[]𝐴𝑄𝐪𝐵𝑄𝐪𝐵𝑄𝐪𝐴𝑄𝐪\displaystyle\times\left[\begin{array}[]{cc}A(-Q;{\bf q})&B(Q;{\bf q})\\ B(Q;{\bf q})&A(Q;{\bf q})\end{array}\right]× [ start_ARRAY start_ROW start_CELL italic_A ( - italic_Q ; bold_q ) end_CELL start_CELL italic_B ( italic_Q ; bold_q ) end_CELL end_ROW start_ROW start_CELL italic_B ( italic_Q ; bold_q ) end_CELL start_CELL italic_A ( italic_Q ; bold_q ) end_CELL end_ROW end_ARRAY ] (34)

where

A(Q;𝐪)A𝑄𝐪\displaystyle\mathrm{A}(Q;\mathbf{q})roman_A ( italic_Q ; bold_q ) =m4πaF+d𝐤(2π)3m𝐤2absent𝑚4𝜋subscript𝑎𝐹𝑑𝐤superscript2𝜋3𝑚superscript𝐤2\displaystyle=-\frac{m}{4\pi a_{F}}+\int\frac{d\mathbf{k}}{(2\pi)^{3}}\frac{m}% {\mathbf{k}^{2}}= - divide start_ARG italic_m end_ARG start_ARG 4 italic_π italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_ARG + ∫ divide start_ARG italic_d bold_k end_ARG start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG divide start_ARG italic_m end_ARG start_ARG bold_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG
k𝒢11mf(k+Q;𝐪)𝒢11mf(k;𝐪)subscript𝑘superscriptsubscript𝒢11mf𝑘𝑄𝐪superscriptsubscript𝒢11mf𝑘𝐪\displaystyle-\sum_{k}\mathcal{G}_{11}^{\mathrm{mf}}(k+Q;\mathbf{q})\mathcal{G% }_{11}^{\mathrm{mf}}(-k;\mathbf{q})- ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_mf end_POSTSUPERSCRIPT ( italic_k + italic_Q ; bold_q ) caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_mf end_POSTSUPERSCRIPT ( - italic_k ; bold_q ) (35)
B(Q;𝐪)𝐵𝑄𝐪\displaystyle B(Q;\mathbf{q})italic_B ( italic_Q ; bold_q ) =k𝒢12mf(k+Q;𝐪)𝒢12mf(k;𝐪)absentsubscript𝑘superscriptsubscript𝒢12mf𝑘𝑄𝐪superscriptsubscript𝒢12mf𝑘𝐪\displaystyle=\sum_{k}\mathcal{G}_{12}^{\mathrm{mf}}(k+Q;\mathbf{q})\mathcal{G% }_{12}^{\mathrm{mf}}(-k;\mathbf{q})= ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT caligraphic_G start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_mf end_POSTSUPERSCRIPT ( italic_k + italic_Q ; bold_q ) caligraphic_G start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_mf end_POSTSUPERSCRIPT ( - italic_k ; bold_q ) (36)

with the normal 𝒢11mf(k;𝐪)superscriptsubscript𝒢11mf𝑘𝐪\mathcal{G}_{11}^{\mathrm{mf}}(k;{\bf q})caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_mf end_POSTSUPERSCRIPT ( italic_k ; bold_q ) and anomalous 𝒢12mf(k;𝐪)superscriptsubscript𝒢12mf𝑘𝐪\mathcal{G}_{12}^{\mathrm{mf}}(k;{\bf q})caligraphic_G start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_mf end_POSTSUPERSCRIPT ( italic_k ; bold_q ) mean-field Green’s functions given by Eqs. (A). Performing the sums over the Matsubara frequency in Eqs. (35) and (36), one obtains for the particle-particle rungs the rather lengthy expressions reported in detail in Appendix A of Ref. Pisani-2023 , which need not be reported here. It may be instead important to emphasize that the presence of the pair propagator (31) in the t𝑡titalic_t-matrix approach guarantees that the effects not only of pair-breaking (single-particle) excitations but also of sound-mode (two-particle) excitations are present in the physical quantities calculated in terms of this approach.

In addition, in Appendix B of Ref. Pisani-2023 it was shown that, when approximating the quantities A(Q;𝐪)𝐴𝑄𝐪A(Q;{\bf q})italic_A ( italic_Q ; bold_q ) and B(Q;𝐪)𝐵𝑄𝐪B(Q;{\bf q})italic_B ( italic_Q ; bold_q ) in the BEC limit of the BCS-BEC crossover, the expression of the current, obtained by utilizing 𝒢11pf(k;𝐪)superscriptsubscript𝒢11pf𝑘𝐪\mathcal{G}_{11}^{\mathrm{pf}}(k;\mathbf{q})caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_pf end_POSTSUPERSCRIPT ( italic_k ; bold_q ) of Eq. (28) into Eq. (18b), recovers the typical form of the current within a two-fluid model PS-2008 for a bosonic gas treated with the Bogoliubov approximation. This form, in turn, coincides with the expression (10) considered in Sec. II.2.

2. The mLPDA approach

The mLPDA approach introduced in Ref. Pisani-2023 consists of a “modified” Local Phase Density Approximation approach, in which the inclusion of pairing fluctuations at the level of the non-self-consistent t𝑡titalic_t-matrix was implemented on top of the original LPDA approach of Ref. Simonucci-2014 . In particular, the LPDA approach was utilized to study the Josephson effect occurring when a barrier is embedded in a homogeneous fermionic superfluid spanning the BCS-BEC crossover Piselli-2020 , a problem for which the mLPDA approach was further considered Piselli-2023 aiming at comparing with recent experimental measurements of the Josephson critical current in ultra-cold atomic Fermi gases Kwon-2020 ; DelPace-2021 .

When a super-current flows across a barrier embedded in an otherwise homogeneous superfluid, the order parameter reads

Δ(x)=|Δ~(x)|e2i𝐪𝐱+2iϕ(x)=e2i𝐪𝐱Δ~(x),Δ𝑥~Δ𝑥superscript𝑒2𝑖𝐪𝐱2𝑖italic-ϕ𝑥superscript𝑒2𝑖𝐪𝐱~Δ𝑥\Delta(x)=|\tilde{\Delta}(x)|e^{2i\mathbf{q}\cdot\mathbf{x}+2i\phi(x)}=e^{2i% \mathbf{q}\cdot\mathbf{x}}\tilde{\Delta}(x)\,,roman_Δ ( italic_x ) = | over~ start_ARG roman_Δ end_ARG ( italic_x ) | italic_e start_POSTSUPERSCRIPT 2 italic_i bold_q ⋅ bold_x + 2 italic_i italic_ϕ ( italic_x ) end_POSTSUPERSCRIPT = italic_e start_POSTSUPERSCRIPT 2 italic_i bold_q ⋅ bold_x end_POSTSUPERSCRIPT over~ start_ARG roman_Δ end_ARG ( italic_x ) , (37)

where 𝐱^^𝐱\hat{\mathbf{x}}over^ start_ARG bold_x end_ARG is the direction of the superfluid flow and 2ϕ(x)2italic-ϕ𝑥2\phi(x)2 italic_ϕ ( italic_x ) the phase of the order parameter due to the presence of the barrier, in addition to that occurring in Eq. (15).

In this case, the LPDA equation takes the form Piselli-2020

m4πaFΔ~(x)=0(𝐱)Δ~(x)+1(x)4md2dx2Δ~(x)+i1(x)qmdΔ~(x)dx,𝑚4𝜋subscript𝑎𝐹~Δ𝑥subscript0𝐱~Δ𝑥subscript1𝑥4𝑚superscriptd2dsuperscript𝑥2~Δ𝑥𝑖subscript1𝑥𝑞𝑚d~Δ𝑥d𝑥\begin{split}-\dfrac{m}{4\pi a_{F}}\tilde{\Delta}(x)&=\mathcal{I}_{0}(\mathbf{% x})\tilde{\Delta}(x)+\dfrac{\mathcal{I}_{1}(x)}{4m}\dfrac{\text{d}^{2}}{\text{% d}x^{2}}\tilde{\Delta}(x)\\ &+i\mathcal{I}_{1}(x)\dfrac{q}{m}\dfrac{\text{d}\tilde{\Delta}(x)}{\text{d}x},% \end{split}start_ROW start_CELL - divide start_ARG italic_m end_ARG start_ARG 4 italic_π italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_ARG over~ start_ARG roman_Δ end_ARG ( italic_x ) end_CELL start_CELL = caligraphic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_x ) over~ start_ARG roman_Δ end_ARG ( italic_x ) + divide start_ARG caligraphic_I start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_x ) end_ARG start_ARG 4 italic_m end_ARG divide start_ARG d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG d italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG over~ start_ARG roman_Δ end_ARG ( italic_x ) end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL + italic_i caligraphic_I start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_x ) divide start_ARG italic_q end_ARG start_ARG italic_m end_ARG divide start_ARG d over~ start_ARG roman_Δ end_ARG ( italic_x ) end_ARG start_ARG d italic_x end_ARG , end_CELL end_ROW (38)

where the coefficients 0subscript0\mathcal{I}_{0}caligraphic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and 1subscript1\mathcal{I}_{1}caligraphic_I start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT are given by

0(x)=d𝐤(2π)3[12fF(E+𝐪(𝐤|x))2E(𝐤|x)m𝐤2]subscript0𝑥d𝐤superscript2𝜋3delimited-[]12subscript𝑓𝐹superscriptsubscript𝐸𝐪conditional𝐤𝑥2𝐸conditional𝐤𝑥𝑚superscript𝐤2\mathcal{I}_{0}(x)=\int\!\!\dfrac{\text{d}\mathbf{k}}{(2\pi)^{3}}\left[\dfrac{% 1-2f_{F}(E_{+}^{\mathbf{q}}(\mathbf{k}|x))}{2E(\mathbf{k}|x)}-\dfrac{m}{% \mathbf{k}^{2}}\right]caligraphic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x ) = ∫ divide start_ARG d bold_k end_ARG start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG [ divide start_ARG 1 - 2 italic_f start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_E start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT bold_q end_POSTSUPERSCRIPT ( bold_k | italic_x ) ) end_ARG start_ARG 2 italic_E ( bold_k | italic_x ) end_ARG - divide start_ARG italic_m end_ARG start_ARG bold_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ] (39a)
1(x)=12d𝐤(2π)3{ξ(𝐤|x)2E(𝐤|x)3[12fF(E+𝐪(𝐤|x))]+ξ(𝐤|x)2E(𝐤|x)2fF(E+𝐪(𝐤|x))E+𝐪(𝐤|x)+𝐤𝐪𝐪21E(𝐤|x)fF(E+𝐐𝟎(𝐤|x))E+𝐪(𝐤|x)}subscript1𝑥12d𝐤superscript2𝜋3𝜉conditional𝐤𝑥2𝐸superscriptconditional𝐤𝑥3delimited-[]12subscript𝑓𝐹superscriptsubscript𝐸𝐪|𝐤𝑥𝜉conditional𝐤𝑥2𝐸superscriptconditional𝐤𝑥2subscript𝑓𝐹superscriptsubscript𝐸𝐪conditional𝐤𝑥superscriptsubscript𝐸𝐪conditional𝐤𝑥𝐤𝐪superscript𝐪21𝐸conditional𝐤𝑥subscript𝑓𝐹superscriptsubscript𝐸subscript𝐐0conditional𝐤𝑥superscriptsubscript𝐸𝐪conditional𝐤𝑥\begin{split}\mathcal{I}_{1}(x)&=\dfrac{1}{2}\int\!\!\dfrac{\text{d}\mathbf{k}% }{(2\pi)^{3}}\bigg{\{}\dfrac{\xi(\mathbf{k}|x)}{2E(\mathbf{k}|x)^{3}}[1-2f_{F}% (E_{+}^{\mathbf{q}}(\mathbf{k}|x))]\\ &+\dfrac{\xi(\mathbf{k}|x)}{2E(\mathbf{k}|x)^{2}}\dfrac{\partial f_{F}(E_{+}^{% \mathbf{q}}(\mathbf{k}|x))}{\partial E_{+}^{\mathbf{q}}(\mathbf{k}|x)}\\ &+\dfrac{\mathbf{k}\cdot\mathbf{q}}{\mathbf{q}^{2}}\dfrac{1}{E(\mathbf{k}|x)}% \dfrac{\partial f_{F}(E_{+}^{\mathbf{Q_{0}}}(\mathbf{k}|x))}{\partial E_{+}^{% \mathbf{q}}(\mathbf{k}|x)}\bigg{\}}\end{split}start_ROW start_CELL caligraphic_I start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_x ) end_CELL start_CELL = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∫ divide start_ARG d bold_k end_ARG start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG { divide start_ARG italic_ξ ( bold_k | italic_x ) end_ARG start_ARG 2 italic_E ( bold_k | italic_x ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG [ 1 - 2 italic_f start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_E start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT bold_q end_POSTSUPERSCRIPT ( bold_k | italic_x ) ) ] end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL + divide start_ARG italic_ξ ( bold_k | italic_x ) end_ARG start_ARG 2 italic_E ( bold_k | italic_x ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG divide start_ARG ∂ italic_f start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_E start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT bold_q end_POSTSUPERSCRIPT ( bold_k | italic_x ) ) end_ARG start_ARG ∂ italic_E start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT bold_q end_POSTSUPERSCRIPT ( bold_k | italic_x ) end_ARG end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL + divide start_ARG bold_k ⋅ bold_q end_ARG start_ARG bold_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG divide start_ARG 1 end_ARG start_ARG italic_E ( bold_k | italic_x ) end_ARG divide start_ARG ∂ italic_f start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_E start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT bold_Q start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ( bold_k | italic_x ) ) end_ARG start_ARG ∂ italic_E start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT bold_q end_POSTSUPERSCRIPT ( bold_k | italic_x ) end_ARG } end_CELL end_ROW (39b)

with

ξ(𝐤|x)=𝐤22m[μVext(x)𝐪22m]E(𝐤|x)=ξ(𝐤|x)2+|Δ~(x)|2E+𝐪(𝐤|x)=E(𝐤|x)+𝐤𝐪m,𝜉conditional𝐤𝑥superscript𝐤22𝑚delimited-[]𝜇subscript𝑉ext𝑥superscript𝐪22𝑚𝐸conditional𝐤𝑥𝜉superscriptconditional𝐤𝑥2superscript~Δ𝑥2superscriptsubscript𝐸𝐪conditional𝐤𝑥𝐸conditional𝐤𝑥𝐤𝐪𝑚\begin{split}\xi(\mathbf{k}|x)&=\dfrac{\mathbf{k}^{2}}{2m}-\left[\mu-V_{% \mathrm{ext}}(x)-\dfrac{\mathbf{q}^{2}}{2m}\right]\\ E(\mathbf{k}|x)&=\sqrt{\xi(\mathbf{k}|x)^{2}+|\tilde{\Delta}(x)|^{2}}\\ E_{+}^{\mathbf{q}}(\mathbf{k}|x)&=E(\mathbf{k}|x)+\dfrac{\mathbf{k}\cdot% \mathbf{q}}{m}\,,\end{split}start_ROW start_CELL italic_ξ ( bold_k | italic_x ) end_CELL start_CELL = divide start_ARG bold_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG - [ italic_μ - italic_V start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT ( italic_x ) - divide start_ARG bold_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG ] end_CELL end_ROW start_ROW start_CELL italic_E ( bold_k | italic_x ) end_CELL start_CELL = square-root start_ARG italic_ξ ( bold_k | italic_x ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + | over~ start_ARG roman_Δ end_ARG ( italic_x ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_CELL end_ROW start_ROW start_CELL italic_E start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT bold_q end_POSTSUPERSCRIPT ( bold_k | italic_x ) end_CELL start_CELL = italic_E ( bold_k | italic_x ) + divide start_ARG bold_k ⋅ bold_q end_ARG start_ARG italic_m end_ARG , end_CELL end_ROW (40)

Vext(x)subscript𝑉ext𝑥V_{\mathrm{ext}}(x)italic_V start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT ( italic_x ) being the external potential associated with the Josephson barrier.

When implementing the numerical calculations, the imaginary part of the LPDA equation (38) is conveniently replaced by the constraint of the current conservation Piselli-2020 ; Pisani-2023 ; Piselli-2023 , namely,

j(x)J=0,𝑗𝑥𝐽0j(x)-J=0\,,italic_j ( italic_x ) - italic_J = 0 , (41)

where J𝐽Jitalic_J is the current evaluated far from the barrier.

Within the LPDA approach of Ref. Simonucci-2014 , the expression for the local current j(x)𝑗𝑥j(x)italic_j ( italic_x ) to be utilized in Eq. (41) is obtained by entering the mean-filed form (20a) for 𝒢11(k;𝐪)subscript𝒢11𝑘𝐪\mathcal{G}_{11}\left(k;\mathbf{q}\right)caligraphic_G start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( italic_k ; bold_q ) into the expression (18b) for the current, provided the following local replacements are performed therein:

μμVext(x)𝜇𝜇subscript𝑉ext𝑥\mu\longrightarrow\mu-V_{\mathrm{ext}}(x)italic_μ ⟶ italic_μ - italic_V start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT ( italic_x ) (42a)
Δ𝐪|Δ(x)|subscriptΔ𝐪Δ𝑥\Delta_{\mathbf{q}}\longrightarrow|\Delta(x)|roman_Δ start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT ⟶ | roman_Δ ( italic_x ) | (42b)
𝐪𝐪+dϕ(x)dx.𝐪𝐪𝑑italic-ϕ𝑥𝑑𝑥\mathbf{q}\longrightarrow\mathbf{q}+\dfrac{d\phi(x)}{dx}\,.bold_q ⟶ bold_q + divide start_ARG italic_d italic_ϕ ( italic_x ) end_ARG start_ARG italic_d italic_x end_ARG . (42c)

Within the mLPDA approach, on the other hand, the local replacement (42a) may lead to unwanted unphysical singularities in the diagonal element Γ11(Q;𝐪)subscriptΓ11𝑄𝐪\Gamma_{11}(Q;\mathbf{q})roman_Γ start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ( italic_Q ; bold_q ) (31) of the pair propagator, when this is treated within a local perspective with local values of the gap parameter and of the chemical potential. For this reason, in Ref. Pisani-2023 it was found convenient to utilize the local requirement

μμVeff(x)𝜇𝜇subscript𝑉eff𝑥\mu\longrightarrow\mu-V_{\mathrm{eff}}(x)italic_μ ⟶ italic_μ - italic_V start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ( italic_x ) (43)

in the place of Eq. (42a), where now Veff(x)subscript𝑉eff𝑥V_{\mathrm{eff}}(x)italic_V start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ( italic_x ) is a suitable “effective” potential which ensures the gapless condition at Q=0𝑄0Q=0italic_Q = 0 of the pair propagator at any x𝑥xitalic_x. Examples of the spatial profile of Veff(x)subscript𝑉eff𝑥V_{\mathrm{eff}}(x)italic_V start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ( italic_x ) were reported in Fig. 2 of Ref. Pisani-2023 for several couplings and temperatures, when the external potential Veff(x)subscript𝑉eff𝑥V_{\mathrm{eff}}(x)italic_V start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ( italic_x ) has a Gaussian form.

3. The extended Gorkov-Melik-Barkhudarov approach

In Ref. Pisani-2018-b a diagrammatic scheme was implemented for improving on the treatment of pairing fluctuations over and above the t𝑡titalic_t-matrix approach, by generalizing to the whole BCS-BEC crossover the original work by Gor’kov and Melik-Barkhudarov (GMB) which was meant for the BCS limit only GMB-1961 . To this end, the property of the pair propagator 𝚪(𝐐,Ω)𝚪𝐐Ω\mathbf{\Gamma}({\bf Q},\Omega)bold_Γ ( bold_Q , roman_Ω ) to be equivalent to the Bogoliubov propagators for composite bosons that form in the BEC limit of the BCS-BEC crossover Andrenacci-2003 was first utilized to show that, already at the level of the t𝑡titalic_t-matrix approach in the superfluid phase, the gap equation for the constituent fermions throughout the whole BCS-BEC crossover is equivalent to a bosonic Hugenholtz-Pines condition, in the form Γ111(0,0)Γ121(0,0)=0superscriptsubscriptΓ11100subscriptsuperscriptΓ112000\Gamma_{11}^{-1}(0,0)-\Gamma^{-1}_{12}(0,0)=0roman_Γ start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( 0 , 0 ) - roman_Γ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT ( 0 , 0 ) = 0. In this way, it was possible to go beyond the t𝑡titalic_t-matrix approach in a natural way, by introducing a suitable bosonic-like self-energy correction 𝚺𝐁(𝐐,Ων)subscript𝚺𝐁𝐐subscriptΩ𝜈\mathbf{\Sigma_{B}}({\bf Q},\Omega_{\nu})bold_Σ start_POSTSUBSCRIPT bold_B end_POSTSUBSCRIPT ( bold_Q , roman_Ω start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ) to the bare pair propagator 𝚪(𝐐,Ων)𝚪𝐐subscriptΩ𝜈\mathbf{\Gamma}({\bf Q},\Omega_{\nu})bold_Γ ( bold_Q , roman_Ω start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ), such that

𝚪1(𝐐,Ων)superscript𝚪1𝐐subscriptΩ𝜈\displaystyle\mathbf{\Gamma}^{-1}({\bf Q},\Omega_{\nu})bold_Γ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( bold_Q , roman_Ω start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ) \displaystyle\rightarrow 𝚪dressed1(𝐐,Ων)superscriptsubscript𝚪dressed1𝐐subscriptΩ𝜈\displaystyle\mathbf{\Gamma}_{\mathrm{dressed}}^{-1}({\bf Q},\Omega_{\nu})bold_Γ start_POSTSUBSCRIPT roman_dressed end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( bold_Q , roman_Ω start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ) (44)
=\displaystyle== 𝚪1(𝐐,Ων)𝚺𝐁(𝐐,Ων)superscript𝚪1𝐐subscriptΩ𝜈subscript𝚺𝐁𝐐subscriptΩ𝜈\displaystyle\mathbf{\Gamma}^{-1}({\bf Q},\Omega_{\nu})-\mathbf{\Sigma_{B}}({% \bf Q},\Omega_{\nu})bold_Γ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( bold_Q , roman_Ω start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ) - bold_Σ start_POSTSUBSCRIPT bold_B end_POSTSUBSCRIPT ( bold_Q , roman_Ω start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT )

in a way formally equivalent to a Dyson’s equation. In addition, in Ref. Pisani-2018-b the fermionic self-energy was kept of the t𝑡titalic_t-matrix form (that is, like in Eq. (27)), with the bare pair propagator replaced, however, by the dressed one of Eq. (44).

As mentioned above, the form of the bosonic-like self-energy 𝚺𝐁subscript𝚺𝐁\mathbf{\Sigma_{B}}bold_Σ start_POSTSUBSCRIPT bold_B end_POSTSUBSCRIPT adopted in Ref. Pisani-2018-b was motivated by the original work by Gor’kov and Melik-Barkhudarov GMB-1961 , who considered the screening of the pairing interaction due to the polarization of the surrounding medium, as represented diagrammatically at second order in the inter-particle interaction by two crossing interaction lines in a particle-hole rung. This effect was shown to introduce a correction by a factor of (4e)1/3superscript4𝑒13(4e)^{1/3}( 4 italic_e ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT to the BCS values of both the critical temperature and the pairing gap GMB-1961 . In Ref. Pisani-2018-b the GMB correction of Ref. GMB-1961 was extended to the whole BCS-BEC crossover, both in the normal and superfluid phases, by identifying the diagram representing the bosonic-like self-energy to be inserted in Eq. (44) with the original GMB diagram, where the interaction lines of the particle-hole rung are now replaced by pair propagators 𝚪(𝐐,Ω)𝚪𝐐Ω\mathbf{\Gamma}({\bf Q},\Omega)bold_Γ ( bold_Q , roman_Ω ) which include infinitely repeated scattering between two fermions and thus go beyond second order (cf. Fig. 2 of Ref. Pisani-2018-b ).

In Ref. Pisani-2018-b the fermionic superfluid was considered at rest, which corresponds to the case with 𝐪=0𝐪0\mathbf{q}=0bold_q = 0 in Eqs. (27)-(36). Full consideration of the GMB correction in the presence of a superfluid flow with 𝐪0𝐪0\mathbf{q}\neq 0bold_q ≠ 0 is beyond the scope of the present work. Nonetheless, we may still consider the effect of the GMB correction in an approximate manner, by replacing the bare pair propagator 𝚪𝚪\mathbf{\Gamma}bold_Γ of Eq. (44) by the dressed one 𝚪dressedsubscript𝚪dressed\mathbf{\Gamma}_{\mathrm{dressed}}bold_Γ start_POSTSUBSCRIPT roman_dressed end_POSTSUBSCRIPT of Eq. (44), where now 𝚪(𝐐,Ων)𝚪(𝐐,Ων;𝐪)𝚪𝐐subscriptΩ𝜈𝚪𝐐subscriptΩ𝜈𝐪\mathbf{\Gamma}({\bf Q},\Omega_{\nu})\rightarrow\mathbf{\Gamma}({\bf Q},\Omega% _{\nu};{\bf q})bold_Γ ( bold_Q , roman_Ω start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ) → bold_Γ ( bold_Q , roman_Ω start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ; bold_q ) contains 𝐪0𝐪0\mathbf{q}\neq 0bold_q ≠ 0 while the bosonic-like self-energy correction 𝚺𝐁subscript𝚺𝐁\mathbf{\Sigma_{B}}bold_Σ start_POSTSUBSCRIPT bold_B end_POSTSUBSCRIPT is taken from the expression with 𝐪=0𝐪0\mathbf{q}=0bold_q = 0 given in Ref. Pisani-2018-b in the absence of current, where we further set 𝐐=0𝐐0{\bf Q}=0bold_Q = 0 and Ων=0subscriptΩ𝜈0\Omega_{\nu}=0roman_Ω start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT = 0. As already noted in Ref. Pisani-2018-b , at low temperature the effect of the GMB self-energy amounts in practice to a coupling-dependent shift of the coupling strength (kFaF)1superscriptsubscript𝑘𝐹subscript𝑎𝐹1(k_{F}a_{F})^{-1}( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT entering in the bare inverse propagator 𝚪1(𝐐,Ων)superscript𝚪1𝐐subscriptΩ𝜈\mathbf{\Gamma}^{-1}({\bf Q},\Omega_{\nu})bold_Γ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( bold_Q , roman_Ω start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ). This remark allows for a swift implementation of the GMB correction in the present work, as it was done in Sec. III-C.

Appendix B ADDENDUM ABOUT TWO TOPICS TREATED IN THE MAIN TEXT

This Appendix expands on two topics dealt with in Secs. II.1 and III.2, regarding respectively: (i) The free energy of a weakly-attractive Fermi gas treated at the mean-field level, thus illustrating in more detail the Bardeen criterion for the supercritical flow (given by Eq. (4) of Sec. II.1) in terms of an analytically solvable approximation; (ii) The crossing between ascending and descending branches of two-particle excitations, which is shown to occur on the BEC side of unitarity for a coupling at which the underlying Fermi surface has not yet collapsed.

1. Free energy at the mean-field level in the presence of a super-current

Refer to caption
Figure 9: Upper panel: The free energy density Fs(q)subscript𝐹𝑠𝑞F_{s}(q)italic_F start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_q ) given by Eq. (B) (red full line) is shown vs the magnitude q𝑞qitalic_q of the momentum (in units of the Fermi wave vector kFsubscript𝑘𝐹k_{F}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT) for coupling (kFaF)1=1.0superscriptsubscript𝑘𝐹subscript𝑎𝐹11.0(k_{F}a_{F})^{-1}=-1.0( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = - 1.0, where it is also compared with the quantity Fs(q=0)+q22msubscript𝐹𝑠𝑞0superscript𝑞22𝑚F_{s}(q=0)+\frac{q^{2}}{2m}italic_F start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_q = 0 ) + divide start_ARG italic_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG (brown dashed line) and with the free energy density Fn=35nEFsubscript𝐹𝑛35𝑛subscript𝐸𝐹F_{n}={3\over 5}nE_{F}italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = divide start_ARG 3 end_ARG start_ARG 5 end_ARG italic_n italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT of an ideal Fermi gas (horizontal black dotted line). Also shown are the Landau critical value qLsubscript𝑞Lq_{\mathrm{L}}italic_q start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT (vertical green dashed-dotted line) and the Bardeen critical value qBsubscript𝑞Bq_{\mathrm{B}}italic_q start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT (vertical blue dashed line). The inset shows the difference Fs(q=0)+q22mFs(q)subscript𝐹𝑠𝑞0superscript𝑞22𝑚subscript𝐹𝑠𝑞F_{s}(q=0)+\frac{q^{2}}{2m}-F_{s}(q)italic_F start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_q = 0 ) + divide start_ARG italic_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG - italic_F start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_q ) in the range between (about) qLsubscript𝑞Lq_{\mathrm{L}}italic_q start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT and qBsubscript𝑞Bq_{\mathrm{B}}italic_q start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT. Lower panel: The corresponding superfluid order parameter ΔqsubscriptΔ𝑞\Delta_{q}roman_Δ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT (orange dashed line) and superfluid density ρs(q)subscript𝜌𝑠𝑞\rho_{s}(q)italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_q ) (violet full line) are shown vs q𝑞qitalic_q.

The free energy density Fs(𝐪)subscript𝐹𝑠𝐪F_{s}({\bf q})italic_F start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( bold_q ) of a weakly-attractive Fermi gas treated at the mean-field level can be computed in the superfluid phase at the mean-field level by following Ref. Pistolesi-1996 , provided that the Nambu Green’s functions in Eq. (2.35) therein are replaced by their counterpart (A) in the presence of a super-current. In this way, one obtains the following expression

Fs(𝐪)subscript𝐹𝑠𝐪\displaystyle F_{s}({\bf q})italic_F start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( bold_q ) =\displaystyle== Δq2v0+d𝐤(2π)3(ξ(𝐤,𝐪)E(𝐤,𝐪))superscriptsubscriptΔ𝑞2subscript𝑣0𝑑𝐤superscript2𝜋3𝜉𝐤𝐪𝐸𝐤𝐪\displaystyle-{\Delta_{q}^{2}\over v_{0}}+\int\frac{d{\bf k}}{(2\pi)^{3}}\bigg% {(}\xi({\bf k},{\bf q})-E({\bf k},{\bf q})\bigg{)}- divide start_ARG roman_Δ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG + ∫ divide start_ARG italic_d bold_k end_ARG start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG ( italic_ξ ( bold_k , bold_q ) - italic_E ( bold_k , bold_q ) )
\displaystyle-- 2βd𝐤(2π)3log[1+exp(βE+(𝐤,𝐪))]+μ𝐪n2𝛽𝑑𝐤superscript2𝜋31𝛽subscript𝐸𝐤𝐪subscript𝜇𝐪𝑛\displaystyle{2\over\beta}\int\frac{d{\bf k}}{(2\pi)^{3}}\log\left[1+\exp\bigg% {(}\!\!\!-\beta E_{+}({\bf k},{\bf q})\!\bigg{)}\right]+\mu_{{\bf q}}ndivide start_ARG 2 end_ARG start_ARG italic_β end_ARG ∫ divide start_ARG italic_d bold_k end_ARG start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG roman_log [ 1 + roman_exp ( - italic_β italic_E start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ( bold_k , bold_q ) ) ] + italic_μ start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT italic_n

with the notation (22) (cf. also Eq. (42) of Ref. Taylor-2006 ). In the above expression, v0<0subscript𝑣00v_{0}<0italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT < 0 is the strength of the contact inter-particle interaction and μ𝐪subscript𝜇𝐪\mu_{{\bf q}}italic_μ start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT is the chemical potential in the presence of a super-current which depends on q=|𝐪|𝑞𝐪q=|{\bf q}|italic_q = | bold_q | when q>qL𝑞subscript𝑞Lq>q_{\mathrm{L}}italic_q > italic_q start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT.

Limiting to zero temperature, the quantity (B) for coupling (kFaF)1=1.0superscriptsubscript𝑘𝐹subscript𝑎𝐹11.0(k_{F}a_{F})^{-1}=-1.0( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = - 1.0 is shown vs q=|𝐪|𝑞𝐪q=|{\bf q}|italic_q = | bold_q | in the upper panel of Fig. 9 (red full line), where it is compared with its counterpart in the normal state Fn=35nEFsubscript𝐹𝑛35𝑛subscript𝐸𝐹F_{n}={3\over 5}nE_{F}italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = divide start_ARG 3 end_ARG start_ARG 5 end_ARG italic_n italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT with Fermi energy EF=kF22msubscript𝐸𝐹superscriptsubscript𝑘𝐹22𝑚E_{F}=\frac{k_{F}^{2}}{2m}italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = divide start_ARG italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG (horizontal black dotted line) which represents the free energy density of an ideal Fermi gas. This panel shows also the quantity Fs(q=0)+q22msubscript𝐹𝑠𝑞0superscript𝑞22𝑚F_{s}(q=0)+\frac{q^{2}}{2m}italic_F start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_q = 0 ) + divide start_ARG italic_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG (brown dashed line), which benchmarks the superfluid non-dissipative state and thus coincides with Fs(𝐪)subscript𝐹𝑠𝐪F_{s}({\bf q})italic_F start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( bold_q ) up to the Landau critical value qLsubscript𝑞Lq_{\mathrm{L}}italic_q start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT (vertical green dashed-dotted line) at which dissipation sets in. Also shown is the Bardeen critical value qBsubscript𝑞Bq_{\mathrm{B}}italic_q start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT (vertical blue dashed line) in correspondence with the inset of Fig. 1 of the main text which shows j(q)=Fs(q)/q𝑗𝑞subscript𝐹𝑠𝑞𝑞j(q)=\partial F_{s}(q)/\partial qitalic_j ( italic_q ) = ∂ italic_F start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_q ) / ∂ italic_q. In the inset of Fig. 9, the difference Fs(q=0)+q22mFs(q)subscript𝐹𝑠𝑞0superscript𝑞22𝑚subscript𝐹𝑠𝑞F_{s}(q=0)+\frac{q^{2}}{2m}-F_{s}(q)italic_F start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_q = 0 ) + divide start_ARG italic_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG - italic_F start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_q ) is plotted in a restricted range between about qLsubscript𝑞Lq_{\mathrm{L}}italic_q start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT and qBsubscript𝑞Bq_{\mathrm{B}}italic_q start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT, where this quantity starts being different from zero at qLsubscript𝑞Lq_{\mathrm{L}}italic_q start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT.

In the lower panel of Fig. 9 the superfluid order parameter ΔqsubscriptΔ𝑞\Delta_{q}roman_Δ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT of Eq. (15) (orange dashed line) is shown vs q𝑞qitalic_q, where it is seen to vanish way beyond the Bardeen critical value qBsubscript𝑞Bq_{\mathrm{B}}italic_q start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT. Also shown is the superfluid density ρs(q)=Fs2/q2subscript𝜌𝑠𝑞superscriptsubscript𝐹𝑠2superscript𝑞2\rho_{s}(q)=\partial F_{s}^{2}/\partial q^{2}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_q ) = ∂ italic_F start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ∂ italic_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (violet full line) given by Eq. (5) of the main text, which is seen to coincide with the full particle density n𝑛nitalic_n up to the Landau critical value qLsubscript𝑞Lq_{\mathrm{L}}italic_q start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT, after which it abruptly decreases and eventually vanishes at the Bardeen critical value qBsubscript𝑞Bq_{\mathrm{B}}italic_q start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT. At this point, where the superfluid density ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT would change sign, the thermodynamical stability of the system is lost. Since ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT represents the stiffness (or rigidity) of the superfluid component to a perturbing velocity field, its turning negative signals the instability of the translational state of the system, in analogy with the occurrence of a negative compressibility in an unstable mechanical system (like for the liquid-gas transition). The system thus undergoes a phase transition to a new state, which in the present case would most likely involve the condensation of finite-momentum Cooper pairs.

2. Crossing between ascending and descending branches of two-particle excitations

One may wonder whether there exists a connection, between the coupling at which the crossover from pair-breaking to phonon excitations occurs (which determines the maximum current in Figs. 4 and 5) and the coupling at which the underlying Fermi surface collapses (such that at this “splitting point” the single-particle dispersion switches over from BCS-type to bosonic-type Son-2006 ). The detailed analysis reported below shows that this is actually not the case, such that the two couplings (albeit close in values) are physically unrelated to each other.

Refer to caption
Figure 10: Two-particle excitation spectrum for the crossover coupling (kFaF)1=+0.36superscriptsubscript𝑘𝐹subscript𝑎𝐹10.36(k_{F}a_{F})^{-1}=+0.36( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = + 0.36 (red curves) and the weaker coupling (kFaF)1=+0.10superscriptsubscript𝑘𝐹subscript𝑎𝐹10.10(k_{F}a_{F})^{-1}=+0.10( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = + 0.10 (blue curves), where the solid lines correspond to the sound mode and the dashed lines to the onset of the pair-breaking spectrum at zero temperature. The black dotted line represents the Landau critical velocity for (kFaF)1=+0.36superscriptsubscript𝑘𝐹subscript𝑎𝐹10.36(k_{F}a_{F})^{-1}=+0.36( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = + 0.36, where this slope is identical for both pair-breaking and sound mode excitations. [Here, the wave vector Q𝑄Qitalic_Q is in units of the Fermi wave vector kFsubscript𝑘𝐹k_{F}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT and the frequency ΩΩ\Omegaroman_Ω in units of the Fermi energy EFsubscript𝐸𝐹E_{F}italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT.]

In our calculation, the maximum of the Landau velocity equals 0.33vF0.33subscript𝑣𝐹0.33v_{F}0.33 italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT obtained for the coupling (kFaF)1=+0.36superscriptsubscript𝑘𝐹subscript𝑎𝐹10.36(k_{F}a_{F})^{-1}=+0.36( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = + 0.36 (cf. the red full line in Fig. 4). This coupling corresponds to the point where the pair-breaking and sound velocities are comparable to each other (cf. the blue dotted line and the black dashed-dotted line in Fig. 4), such that the two types of excitations have comparable energy. To better illustrate how the change between one type (pair-breaking) to the other type (sound mode) of excitations comes about, following what was done in Fig. 7 of Ref. Pieri-2004 , Fig. 10 shows the zero-temperature two-particle excitation spectrum both for the crossover coupling (kFaF)1=+0.36superscriptsubscript𝑘𝐹subscript𝑎𝐹10.36(k_{F}a_{F})^{-1}=+0.36( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = + 0.36 (red curves) and for the weaker coupling (kFaF)1=+0.10superscriptsubscript𝑘𝐹subscript𝑎𝐹10.10(k_{F}a_{F})^{-1}=+0.10( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = + 0.10 (blue curves), where the solid lines correspond to the sound mode and the dashed lines to the onset of the pair-breaking spectrum. The latter quantity is obtained like in Ref. Pieri-2004 , but now considering the renormalized chemical potential μL=kL2/(2m)subscript𝜇Lsuperscriptsubscript𝑘L22𝑚\mu_{\mathrm{L}}=k_{\mathrm{L}}^{2}/(2m)italic_μ start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT = italic_k start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 2 italic_m ) (where kLsubscript𝑘Lk_{\mathrm{L}}italic_k start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT is the so-called “Luttinger” wave vector to be discussed below) and the excitation gap ΔesubscriptΔe\Delta_{\mathrm{e}}roman_Δ start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT as extracted from the single-particle excitation spectrum, rather than the thermodynamic values μ𝜇\muitalic_μ and ΔΔ\Deltaroman_Δ utilized in Ref. Pieri-2004 (including in this way some degree of self-consistency in the calculation which takes into account the effect of quantum fluctuations. In Fig. 10 the slope of the black dotted line represents the Landau critical velocity for the crossover coupling (kFaF)1=+0.36superscriptsubscript𝑘𝐹subscript𝑎𝐹10.36(k_{F}a_{F})^{-1}=+0.36( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = + 0.36, where this slope is identical for both pair-breaking and sound mode excitations. For the weaker coupling (kFaF)1=+0.10superscriptsubscript𝑘𝐹subscript𝑎𝐹10.10(k_{F}a_{F})^{-1}=+0.10( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = + 0.10, on the other hand, the slope of the sound mode is evidently larger than that of the onset of the pair-breaking excitations, consistently with what is reported in Fig. 4.

Regarding instead the coupling at which the collapse of the underlying Fermi surface takes place upon approaching the BEC side of the BCS-BEC crossover, this coupling refers to a property of the single-particle excitation spectrum of the Fermi system and has been identified as the coupling at which the so-called Luttinger wave vector kLsubscript𝑘𝐿k_{L}italic_k start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT mentioned above vanishes. In particular, this quantity was determined at the critical temperature Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT by including pairing fluctuations beyond mean field within the t𝑡titalic_t-matrix approximation, either when comparing with the available experimental data in ultra-cold Fermi gases Perali-2011 , or when considering in details the properties of the single-particle spectral function in the normal phase of a Fermi gas Palestini-2012 , and even with the inclusion of disorder Palest-2013 . These calculations show that the splitting point “κ0subscript𝜅0\kappa_{0}italic_κ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT” of Ref. Son-2006 , which is the coupling where kLsubscript𝑘𝐿k_{L}italic_k start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT vanishes according to Refs. Perali-2011 ; Palestini-2012 ; Palestini-2013, is about +0.600.60+0.60+ 0.60. This confirms our expectation that the crossover coupling (kFaF)1=+0.36superscriptsubscript𝑘𝐹subscript𝑎𝐹10.36(k_{F}a_{F})^{-1}=+0.36( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = + 0.36 (for two-particle excitations) and the splitting point (for single-particle excitations) are two unrelated quantities.

References

  • (1) M. Tinkham, Introduction to Superconductivity (McGraw-Hill, New York, 1996).
  • (2) E. M. Lifshitz and L. P. Pitaevskii, Statistical Physics, Part 2 (Pergamon Press, Oxford, 1980).
  • (3) P. Fulde, Tunneling density of states for a superconductor carrying a current, Phys. Rev. 137, A783 (1965).
  • (4) R. H. Parmenter, Gapless high-current superconductivity, Phys. Rev. 140, A1952 (1965).
  • (5) A. A. Abrikosov and L. P. Gor’kov, Contribution to the theory of superconducting alloys with paramagnetic impurities, Sov. Phys. JETP 12, 1243 (1961) [Zh. Eksp. Teor. Fiz. 39, 1781 (1960)].
  • (6) K. Maki, in Superconductivity Vol. 2, D. Parks Ed., (M. Dekker, New York, 1969).
  • (7) P. Fulde and R. A. Ferrel, Critical current and electron depairing in superconducting films, Phys. Rev. 131, 2457 (1963).
  • (8) J. Bardeen, Critical fields and currents in superconductors, Rev. Mod. Phys. 34, 667 (1962).
  • (9) M. Yu. Kupriyanov and V. F. Lukichev, Temperature dependence of pair-breaking current in superconductors, Sov. J. Low Temp. Phys. 6, 210 (1980) [Fiz. Nizk. Temp. 6, 445 (1980)].
  • (10) A. L. Fetter, Non uniform states of an imperfect Bose gas, Annals of Physics 70, 67 (1972).
  • (11) L. Kramer, Thermodynamic stability of superfluid flow through a channel, Phys. Rev. 179, 148 (1969).
  • (12) P. Nozières, Theory of Interacting Fermi Systems (Westview Press, USA, 1997).
  • (13) R. P. Feynman and M. Cohen, Energy spectrum of the excitations in Helium, Phys. Rev. 102, 1189 (1956).
  • (14) G. Calvanese Strinati, P. Pieri, G. Röpke, P. Schuck, and M. Urban, The BCS-BEC crossover: From ultra-cold Fermi gases to nuclear systems, Phys. Rep. 738, 1 (2018).
  • (15) M. W. Zwierlein, J. R. Abo-Shaeer, A. Schirotzek, C. H. Schunck, and W. Ketterle, Vortices and superfluidity in a strongly interacting Fermi gas, Nature 435, 1047 (2005).
  • (16) S. Simonucci, P. Pieri, and G. Calvanese Strinati, Vortex arrays in neutral trapped Fermi gases through the BCS–BEC crossover, Nat. Phys. 11, 941 (2015).
  • (17) D. Miller, J. Chin, C. Stan, Y. Liu, W. Setiawan, C. Sanner, and W. Ketterle, Critical velocity for superfluid flow across the BEC-BCS crossover, Phys. Rev. Lett. 99, 070402 (2007).
  • (18) A. Spuntarelli, P. Pieri, and G. Calvanese Strinati, Josephson effect throughout the BCS-BEC crossover, Phys. Rev. Lett. 99, 040401 (2007).
  • (19) M. W. Zwierlein, A. Schirotzek, C. H. Schunck, and W. Ketterle, Fermionic superfluidity with imbalanced spin populations, Science 311, 492 (2006).
  • (20) G. B. Partridge, W. Li, R. I. Kamar, Y-an Liao, and R. G. Hulet, Pairing and phase separation in a polarized Fermi gas, Science 311, 503 (2006).
  • (21) P. Pieri and G. Calvanese Strinati, Trapped fermions with density imbalance in the Bose-Einstein condensate limit, Phys. Rev. Lett. 96, 150404 (2006).
  • (22) W. Weimer, K. Morgener, V. P. Singh, J. Siegl, K. Hueck, N. Luick, L. Mathey, and H. Moritz, Critical velocity in the BEC-BCS crossover, Phys. Rev. Lett. 114, 095301 (2015).
  • (23) G. Del Pace, K. Xhani, A. Muzi Falconi, M. Fedrizzi, N. Grani, D. Hernandez Rajkov, M. Inguscio, F. Scazza, W. J. Kwon, and G. Roati, Imprinting persistent currents in tunable fermionic rings, Phys. Rev. X 12, 041037 (2022).
  • (24) L. D. Landau, The theory of superfluidity of helium II, Zh. Eksp. Teor. Fiz. 11, 592 (1941) [J. Phys. USSR 5, 71 (1941)].
  • (25) L. Pisani, V. Piselli, and G. Calvanese Strinati, Inclusion of pairing fluctuations in the differential equation for the gap parameter for superfluid fermions in the presence of nontrivial spatial constraints, Phys. Rev. B 108, 214503 (2023).
  • (26) P. W. Aderson, Random-Phase Approximation in the theory of superconductivity, Phys. Rev. 112, 1900 (1958).
  • (27) M. Marini, F. Pistolesi, and G. Calvanese Strinati, Evolution from BCS superconductivity to Bose condensation: analytic results for the crossover in three dimensions, Eur. Phys. J. B 1, 151 (1998).
  • (28) A. Spuntarelli, P. Pieri, and G. Calvanese Strinati, Solution of the Bogoliubov-de Gennes equations at zero temperature throughout the BCS-BEC crossover: Josephson and related effects, Phys. Rep. 488, 111 (2010).
  • (29) R. Combescot, M. Yu. Kagan, and S. Stringari, Collective mode of homogeneous superfluid Fermi gases in the BEC-BCS crossover, Phys. Rev. A 74, 042717 (2006).
  • (30) L. Pisani, P. Pieri, and G. Calvanese Strinati, Gap equation with pairing correlations beyond the mean-field approximation and its equivalence to a Hugenholtz-Pines condition for fermion pairs, Phys, Rev. B 98, 104507 (2018).
  • (31) E. Brun Hansen, Infinite conductivity of ordinary and gapless superconductors, Physica 39, 271 (1968).
  • (32) I. M. Khalatnikov, An Introduction to the Theory of Superfluidity (Westview Press, New York, 2000).
  • (33) V. Piselli, S. Simonucci, and G. Calvanese Strinati, Josephson effect at finite temperature along the BCS-BEC crossover, Phys. Rev. B 102, 144517 (2020).
  • (34) H. Godfrin, M. Meschke, H.-J. Lauter, A. Sultan, H. M. Böhm, E. Krotscheck, and M. Panholzer, Observation of a roton collective mode in a two-dimensional Fermi liquid, Nature 483, 576 (2012).
  • (35) L. Chomaz, R. M. W. van Bijnen, D. Petter, G. Faraoni, S. Baier, J. H. Becher, M. J. Mark, F. Wächtler, L. Santos, and F. Ferlaino, Observation of roton mode population in a dipolar quantum gas, Nat Phys. 14, 442 (2018).
  • (36) T. Kubo, Superfluid flow in disordered superconductors with Dynes pair-breaking scattering: Depairing current, kinetic inductance, and superheating field, Phys. Rev. Res. 2, 033203 (2020).
  • (37) M. N. Kunchur, Evaluating superconductors through current induced depairing, Condens. Matter 4, 54 (2019).
  • (38) J. R. Clem and V. G. Kogan, Kinetic impedance and depairing in thin and narrow superconducting films, Phys. Rev. B 86, 174521 (2012).
  • (39) S. V. Iordanskii and L. P. Pitaevskii, Bose condensation of moving rotons, Sov. Phys. Usp. 23, 317 (1980) [Usp. Fiz. Nauk 131, 293 (1980)].
  • (40) I. Carusotto, S. X. Hu, L. A. Collins, and A. Smerzi, Bogoliubov-Cerenkov radiation in a Bose-Einstein condensate flowing against an obstacle, Phys. Rev. Lett. 97, 260403 (2006).
  • (41) G. Baym and C. J. Pethick, Landau critical velocity in weakly interacting Bose gases, Phys. Rev. A 86, 023602 (2012).
  • (42) L. Tanzi, E. Lucioni, F. Famà, J. Catani, A. Fioretti, C. Gabbanini, R. N. Bisset, L. Santos, and G. Modugno, Observation of a dipolar quantum gas with metastable supersolid properties, Phys. Rev. Lett. 122, 130405 (2019).
  • (43) E. Varoquaux, Anderson’s considerations on the flow of superfluid helium: Some offshoots, Rev. Mod. Phys. 87, 803 (2015).
  • (44) P. Pieri, L. Pisani, and G. Calvanese Strinati, BCS-BEC crossover at finite temperature in the broken-symmetry phase, Phys. Rev. B 70, 094508 (2004).
  • (45) H. Biss, L. Sobirey, N. Luick, M. Bohlen, J. J. Kinnunen, G. M. Bruun, T. Lompe, and H. Moritz, Excitation spectrum and superfluid gap of an ultracold Fermi gas, Phys. Rev. Lett. 128, 100401 (2022).
  • (46) M. Link, K. Gao, A. Kell, M. Breyer, D. Eberz, B. Rauf, and M. Köhl, Machine learning the phase diagram of a strongly interacting Fermi gas, Phys. Rev. Lett. 130, 203401 (2023).
  • (47) V. Piselli, L. Pisani, and G. Calvanese Strinati, Josephson current flowing through a nontrivial geometry: Role of pairing fluctuations across the BCS-BEC crossover, Phys. Rev. B 108, 214504 (2023).
  • (48) S. Simonucci and G. Calvanese Strinati, Equation for the superfluid gap obtained by coarse graining the Bogoliubov-de Gennes equations throughout the BCS-BEC crossover, Phys. Rev. B 89, 054511 (2014).
  • (49) F. Palestini and G. Calvanese Strinati, Systematic investigation of the effects of disorder at the lowest order throughout the BCS-BEC crossover, Phys. Rev. B 88, 174504 (2013).
  • (50) P. Pieri and G. Calvanese Strinati, Derivation of the Gross-Pitaevskii equation for condensed bosons from the Bogoliubov-de Gennes equations for superfluid fermions, Phys. Rev. Lett. 91, 030401 (2003).
  • (51) I. V. Brodsky, M. Y. Kagan, A. V. Klaptsov, R. Combescot, and X. Leyronas, Exact diagrammatic approach for dimer-dimer scattering and bound states of three and four resonantly interacting particles, Phys. Rev. A 73, 032724 (2006).
  • (52) V. P. Singh, W. Weimer, K. Morgener, J. Siegl, K. Hueck, N. Luick, H. Moritz, and L. Mathey, Probing superfluidity of Bose-Einstein condensates via laser stirring, Phys. Rev. A 93, 023634 (2016).
  • (53) G. E. Astrakharchik, J. Boronat, J. Casulleras, and S. Giorgini, Equation of state of a Fermi gas in the BEC-BCS crossover: A Quantum Monte Carlo study, Phys. Rev. Lett. 93, 200404 (2004).
  • (54) W. J. Kwon, G. Del Pace, R. Panza, M. Inguscio, W. Zwerger, M. Zaccanti, F. Scazza, and G. Roati, Strongly correlated superfluid order parameters from dc Josephson supercurrents, Science 369, 84 (2020).
  • (55) G. Del Pace, W. J. Kwon, M. Zaccanti, G. Roati, and F. Scazza, Tunneling transport of unitary fermions across the superfluid transition, Phys. Rev. Lett. 126, 055301 (2021).
  • (56) J. File and R. G. Mills, Observation of persistent current in a superconducting solenoid, Phys. Rev. Lett. 10, 93 (1963).
  • (57) K. Xhani, G. Del Pace, F. Scazza, and G. Roati, Decay of persistent currents in annular atomic superfluids, Atoms 11,109 (2023).
  • (58) L. Pisani, A. Perali, P. Pieri, and G. Calvanese Strinati, Entanglement between pairing and screening in the Gorkov-Melik-Barkhudarov correction to the critical temperature throughout the BCS-BEC crossover, Phys. Rev. B 97, 014528 (2018).
  • (59) J. R. Schrieffer, Theory of Superconductivity (Benjamin, New York, 1964).
  • (60) P. G. de Gennes, Superconductivity of Metals and Alloys (Benjamin, New York, 1966).
  • (61) A. L. Fetter and J. D. Walecka, Quantum Theory of Many-Particle Systems (McGraw-Hill, New York, 1971).
  • (62) N. Andrenacci, P. Pieri, and G. Calvanese Strinati, Evolution from BCS superconductivity to Bose-Einstein condensation: Current correlation function in the broken-symmetry phase, Phys. Rev. 68, 144507 (2003).
  • (63) C. J. Pethick and H. Smith, Bose-Einstein Condensation in Dilute Gases (Cambridge Univ. Press, Cambridge, 2008).
  • (64) L. P. Gor’kov and T. M. Melik-Barkhudarov, Contribution to the theory of superfluidity in an imperfect Fermi gas, Sov. Phys. JETP 13,1018 (1961) [Zh. Eksp. Teor. Fiz. (U.S.S.R.) 40, 1452 (1961)].
  • (65) F. Pistolesi and G. Calvanese Strinati, Evolution from BCS superconductivity to Bose condensation: Calculation of the zero-temperature phase coherence length, Phys. Rev. B 53, 15168 (1996).
  • (66) E. Taylor, A. Griffin, N. Fukushima, and Y. Ohashi, Pairing fluctuations and the superfluid density through the BCS-BEC crossover, Phys. Rev. A 74, 063626 (2006).
  • (67) D. T. Son and M. A. Stephanov, Phase diagram of a cold polarized Fermi gas, Phys. Rev. A 74, 013614 (2006).
  • (68) A. Perali, F. Palestini, P. Pieri, G. Calvanese Strinati, J. T. Stewart, J. P. Gaebler, T. E. Drake, and D. S. Jin, Evolution of the normal state of a strongly interacting Fermi gas from a pseudogap phase to a molecular Bose gas, Phys. Rev. Lett. 106, 060402 (2011).
  • (69) F. Palestini, A. Perali, P. Pieri, and G. Calvanese Strinati, Dispersions, weights, and widths of the single-particle spectral function in the normal phase of a Fermi gas, Phys. Rev. B 85, 024517 (2012).